Вы находитесь на странице: 1из 94

Control of an automotive

electromagnetic suspension system

T.P.J. van der Sande


D&C 2011.016

Masters thesis
Coach(es):

ir. B.L.J. Gysen


dr.ir. I.J.M. Besselink

Supervisor:

prof.dr. H. Nijmeijer

Committee:

prof.dr. H. Nijmeijer
dr.ir. I.J.M. Besselink
dr.ir. J.J.H. Paulides

Eindhoven University of Technology


Department of Mechanical Engineering
Master Automotive Technology
Dynamics & Control
Eindhoven, March, 2011

Acknowledgements
First and foremost, my gratitude goes out to ir. Bart Gysen, my direct supervisor for this project.
His ever present support greatly helped me in performing this research.
Second, my thanks go out to prof. Henk Nijmeijer, dr. Igo Besselink, dr. Johan Paulides and
prof. Elena Lomonova. Their guidance, valuable tips and critical questions often helped me in
the right direction.
Thirdly, I would like to thank the EPE group for offering me this interesting and challenging
research topic in which I could further enhance not only my theoretical but also my practical
skills. A special note goes to my roommates, for the interesting discussions.
Finally, I would like to thank my family, girlfriend and other friends for their support and
encouragement. This made my graduation project much more enjoyable.

Abstract
The main research goal of this thesis is to determine what performance gains can be achieved
with a high bandwidth electromagnetic active suspension. As a baseline vehicle a BMW 530i is
used, for which a retrofit electromagnetic suspension consisting of a spring and tubular permanent magnet actuator (TPMA) is designed. To design a control system for this actuator, a model
of the BMW has been created, which consists of a quarter car model with variable sprung mass,
damping coefficient and tire stiffness. As input to this model a road disturbance is used, that was
modeled as a white noise source filtered by a first order low-pass filter. To test the performance
of the actuator and controllers a full size quarter car test setup is used.
As control objectives minimization of the sprung acceleration and dynamic tire compression
are used with constraints on the suspension travel and RMS actuator force. The sprung acceleration is used as an indication for ride comfort and the dynamic tire compression is used as
an indication for handling quality. To account for human sensitivity to vibrations, the ISO2631-1
standard is used to filter the sprung acceleration. The suspension travel of the controlled system
is limited to the maximum value that the BMW achieved with its spring and damper settings over
a given road. Furthermore, the maximum RMS actuator force of 1000 N results from thermal
limits.
Two control approaches are considered, linear quadratic optimal control and robust control.
For the former, a controller is found using a linearized quarter car model. By choosing three
weighting factors either comfort or handling can be emphasized. Variations of the plant are
accounted for by using robust control. Using frequency dependent weighting, certain frequencies
can be emphasized. For instance, human sensitivity to vertical vibrations is incorporated using an
approximation of ISO2631-1. By varying this weighting together with the other weighting filters
either comfort or handling can emphasized, similar to the linear quadratic control.
Measurements on the quarter car setup show that an improvement in comfort of 35 % can be
achieved with linear quadratic control. This differs 55 % from the value predicted by simulations.
However, this deviation can be explained by friction in the test setup and actuator as well as by
uncertainties that were not modeled when designing the LQ controller. In case of the handling
controller, measurements do match the simulations better on the smooth road. Dynamic tire
compression is stability issues of the controller.
With robust control an improvement of 48 % in comfort can be achieved on the setup at the
cost of an increase of 99.3 % in dynamic tire compression. In terms of handling, an improvement of 17.7 % is achieved, worsening comfort by 10.7 %. Frequency weighting clearly has a
desirable effect, as comfort decreases by 6 % for the handling controller on rough road whereas
sprung acceleration worsens by 75 %. This means that all vibrations occur outside of the human
sensitivity range. Deviations of the measurements from the simulations can be explained by stick
slip friction in the suspension actuator as well as vibrations passing through the test setup.

ii

Samenvatting
De belangrijkste onderzoeksvraag van deze thesis is wat voor prestatie winst er kan worden behaald met een hoge bandbreedte electromagnetische actieve ophanging. Als basis voertuig wordt
een BMW 530i gebruikt, waarvoor een retrofit tubulaire permanent magneet actuator, bestaande
uit een veer en actuator, ontworpen is. Om het regelsysteem van deze actuator te ontwerpen is er
een model van de BMW gemaakt dat bestaat uit een kwart voertuig model met variable geveerde
massa, dempings coefficient en band stijfheid. Als ingang voor het model wordt een wegverstoring gebruikt, bestaande uit witte ruis gefilterd met een eerste orde laag doorlaat filter. Om de
prestaties van de actuator en regelaars te bepalen is er een kwart voertuig opstelling gebruikt op
ware grootte.
Het regeldoel is het minimaliseren van de geveerde acceleratie of de dynamische band indrukking met als randvoorwaarden de veerweg en RMS actuator kracht. De afgeveerde versnelling wordt gebruikt om de mate van comfort te bepalen. De dynamische band indrukking
geeft een idee van de kwaliteit van de wegligging. Om rekening te houden met de menselijke
gevoeligheid voor verticale vibraties wordt het ISO2631-1 criterium gebruikt. De limiet op de
veerweg wordt bepaald door de veerweg van de BMW over dezelfde weg, terwijl de 1000 N actuator kracht limiet bepaald wordt door de thermische eigenschappen van deze.
Twee controle topologien worden beschouwd, een linear kwadratisch en robuuste regelaar.
Voor de eerste geldt dat er een optimale regelaar ontworpen wordt aan de hand van een gelinearizeerde versie van het kwart voertuig model. Door het kiezen van drie weegfactoren kunnen
comfort of wegligging benadrukt worden. Om zeker te zijn dat de regelaar stabiel is met de variaties die op kunnen treden in het system wordt een robuuste regeling gebruikt. Deze methode
maakt het mogelijk om frequentie afhankelijke weegfilters te gebruiken. Een voorbeeld hiervan is
het ISO2631-1 criterium, waarvan een benadering van wordt gebruikt om menselijke gevoeligheid
voor verticale vibraties extra te benadrukken. Door dit weegfilter te gebruiken in combinatie met
andere weegfilters kan comfort of wegligging benadrukt worden.
Metingen op de kwart voertuig opstelling laten zien dat comfort met 35 % verbeterd kan
worden met een linear kwadratische regelaar. Dit wijkt 55 % af van de verbetering voorspeld door
simulaties. Dit kan echter verklaard worden door wrijving in de opstelling en actuator alsmede
door onzekerheden die niet meegenomen zijn in het ontwerpen van de LQ regelaar. De resultaten
van de regelaar die ontworpen is voor wegligging komen beter overeen met de simulaties. Een
verbetering van 48.5 % kan worden behaald. Dit kon echter niet worden geverifieerd op de ruwe
weg door instabiliteit van regelaar.
Met de robuuste regelaar kan een verbetering van 48 % worden gehaald in comfort op de
test opstelling ten koste van een verslechtering in dynamische band indrukking van 99.3 %.
Voor wegligging kan er een verbetering van 17.7 % behaald worden waarbij comfort met 10.7 %
verslechterd wordt. De toepassing van frequentie afhankelijke filters heeft een gewenst effect
aangezien comfort met maar 6 % wordt verslechterd terwijl de verticale acceleratie met 75 %
iii

verslechtert. Dit betekent dat alle vibraties optreden buiten het gebied waar mensen het meest
gevoelig zijn. Verschillen tussen de metingen en simulaties kunnen verklaard worden door slickslip wrijving in de actuator en in de test opstelling. Verder spelen vibraties die via het frame van
de test opstelling naar de sensoren komen een rol.

iv

Used Symbols and Abbreviations


Abbreviations
Name
ABC
DOF
LQ
LQG
LQOF
LQR
NS
RC
RS
RP
RMS
TPMA
VAG

Meaning
Active body control
Degrees of freedom
Linear quadratic
Linear quadratic gaussian
Linear quadratic output feedback
Linear quadratic regulator
Nominal stability
Robust control
Robust stability
Robust performance
Root mean square
Tubular permanent magnet actuator
Volkswagen Audi group

Symbols
Symbol

dr
ds
Cr oad
1
Ei
Fact
Fra
Fy
i
J
k Ei
kr
kra
ks
kt
Li
mc
ms
m ra
mu

ns
Q
R
Ri

p
ts
v
Vi
Vx
w
Wi
yc
yr

Meaning
Cut off frequency of road signal
Side slip angle
Lateral tire damping
Sprung damping
Road actuator controller
Uncertainty matrix
Actuator phase back EMF
Suspension actuator force
Road actuator force
Lateral tire force
Current amplitude
LQ control objective
Actuator EMF constant
Lateral tire stiffness
Road actuator spring stiffness
Sprung stiffness
Tire stiffness
Actuator phase inductance
Contact patch mass
Sprung mass
Road actuator mass
Unsprung mass
Structured singular value
Spatial frequency
Weighting matrix for LQ control
Controllability matrix
Actuator phase resistance
Gain that defines road amplitude
Pole pitch
Sampletime
Suspension speed
Supply voltage
Forward velocity
White noise
Weighting filter i
Controlled output
Lateral tire deflection

vi

Symbol
z
zr
zs
zt
zu

Meaning
Suspension travel
Road displacement
Displacement of sprung mass
Tire compression
Displacement of unsprung mass
Speed dependent commutation angle

Conventions
dz (t)
= z (t)
dt

(1)

d 2 z (t)
= z (t)
dt 2

(2)

vii

Contents
1

Introduction
1.1 Problem statement and objectives . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2 Actuator and car model


2.1 BMW 530i . . . . . . . .
2.2 Active suspension system
2.2.1 Actuator . . . . .
2.2.2 Sensors . . . . . .
2.3 Simplified car model . . .
2.3.1 Quarter car model
2.3.2 Road input . . . .
2.4 Summary . . . . . . . . .
3

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

9
9
11
11
16
17
17
18
21

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

22
22
23
25
27
28
30
33
38

4 Analysis of simulation results


4.1 BMW 530i performance on random road .
4.2 Linear quadratic control . . . . . . . . . .
4.3 Robust control . . . . . . . . . . . . . . .
4.4 Summary . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

39
39
42
44
46

.
.
.
.
.

48
48
50
52
53
55

Control of the active suspension


3.1 Control objectives . . . . . . . .
3.2 Duality of control objectives . . .
3.3 Linear quadratic control . . . . .
3.4 Robust control . . . . . . . . . .
3.4.1 Model . . . . . . . . . . .
3.4.2 Robustness requirements
3.4.3 Weighting filters . . . . .
3.5 Summary . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

1
5
5
8

Quarter car test setup


5.1 Description of the test setup . . .
5.2 Control of road actuation . . . .
5.3 Kalman filter suspension travel .
5.4 Experimental validation of setup
5.5 Summary . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

viii

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

6 Measurement results achieved on quarter car setup


6.1 Linear quadratic control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2 Robust control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

57
57
60
67

69
69
71

Conclusions and recommendations


7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

A Tire Model
A.1 Vertical stiffness . . . . .
A.2 Relaxation measurements
A.3 Magic Formula . . . . . .
A.4 Tire parameters . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

B LDIA 2011 digest

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

75
75
77
79
80
81

ix

Chapter 1

Introduction
Test drivers usually emerge from the car with their imagination in overdrive. "The greatest single advance
in car engineering since the war," the British magazine Car declared on the cover of a recent issue. Cars
editor, Steve Cropley, wrote that one could take the benefit of all other modern automobile developments,
"add the up and double the total - and you might come somewhere near the degree to which full active
suspension improves a car." [1]
Although one should always be sceptical about the enthusiasm expressed during such first
time tests, this statement does indicate that active suspension offers the opportunity to change the
performance of a car substantially. Ever since, manufacturers have been hard at work to develop
systems suited for mass production. Examples of this are the Active Body Control (ABC) [2] by
Mercedes, Hydractive [3] from Citron and air suspension used by up market manufacturers to
increase ground clearance in their off-road models and to influence the character of the car (Land
Rover, Audi, VW, Lexus, Lincoln etc.). Next to fully active systems, semi-active systems have also
been developed. Examples are Delphi magneto-rheological dampers [4] used by Ferrari as well as
Cadillac and the VAG group. Alfa Romeo uses a semi active system developed by Magneti Marelli
[5] which controls valves in the damper, thereby changing its characteristics.
It is obvious that numerous suspension suspension systems are already in production, generally they can be divided into three groups: Passive (Figure 1.1(a)), semi-active suspension (Figure
1.1(b)) and active (Figure 1.1(c), (d) and (e)) systems. The main difference between them is that
the former has no possibility of changing the suspension characteristics, whereas the second can
vary the amount of dissipative power. The fully active system can not only vary the amount of
dissipative power, but can also supply power to the system by means of active force generation.
Implementation of the suspension systems is done very differently by various manufacturers.
The Mercedes ABC system for instance, works by means of a hydraulic actuator in series with a
passive spring-damper combination. Its bandwidth, due to valves and connective hoses is only
5 Hz. It is therefore primarily used to level the vehicle. Due to its 200 bar operating pressure
its power demand is in the range of 3-5 kW. Due to this low bandwidth, the suspension becomes
virtually rigid from 10 Hz onwards [6], thereby requiring the passive suspension to provide good
comfort and roadholding beyond that frequency.
The Citron system as shown in Figure 1.2 uses spheres filled with nitrogen and a hydraulic
fluid separated by a rubber membrane to control the ride. When driving over a bump, the fluid
is pushed up the suspension strut compressing the nitrogen and thus providing a spring action.
The hydraulic fluid is then directed through valves, providing damping. When driving normally,
the spheres at the suspension strut are connected to a third sphere increasing the volume of
1

Chassis

Chassis

ks

ds

ks

ds

ds

ks

Chassis

Chassis

Chassis

F
ks

ds

ks

ds

Wheel

Wheel

Wheel

Wheel

Wheel

kt

kt

kt

kt

kt

(a)

(b)

(c)

(d)

(e)

Figure 1.1: Quarter car representation of (a) passive suspension, (b) semi-active suspension, (c) parallel active suspension,
(d) series active suspension and (e) electromagnetic suspension.

the nitrogen and thus providing a lower stiffness and thereby smoother ride. However, when
cornering, valves are closed, disconnecting the central sphere. A firmer ride is achieved this way,
thereby reducing roll of the car. Continuous pressurization of the system is required, making the
power requirement high. A great disadvantage of the system is that when pressure is lost the
vehicle will loose ride height and performance will deteriorate.

Figure 1.2: Citron active suspension system.

The semi-active solutions from Delphi, see Figure 1.3, and Magneti Marelli both influence the
flow of the hydraulic fluid inside the damper. The former uses magneto-rheological fluid, which
changes viscosity when the fluid is exposed to a magnetic field. According to the manufacturer
the damping force is only dependent on the power applied to the magneto-rheological fluid and
can be adjusted up to 1000 times a second. A skyhook control algorithm is used to ensure good
road to wheel contact with the least impulses to the car body. Due to the semi-active nature of
the system, average power is much lower (5 W) compared to the hydraulic suspension systems.
Power can, however, not be supplied to the system, limiting the performance gains of the system
when compared to a passive system.
A novel electro-hydraulic semi-active suspension system is built by Levant Power and is called
the GenShock [7]. It operates by means of a hydraulic cylinder connected to a set of valves and a
2

Figure 1.3: Delphi magneto rheological damper.

hydraulic motor that is connected to a generator. When the vehicle drives over a bump, the linear
motion of the shock absorber pumps the fluid round. The hydraulic motor connected in the same
circuit is then excited by this moving fluid and subsequently excites an electric generator. Electric
energy is then stored in the battery. The manufacturer claims a 1-3% increase in fuel efficiency
and a reduction in vibrations up to 30 %.
An active solution that tries to solve the problem of high power consumption is built by
ZF and Volkswagen [8]. It consists of spindle driven by an electric motor in series with a spring
and in parallel with a conventional damper. By actively controlling the spindle position the series
spring can be loaded, thereby controlling the roll of the vehicle. A skyhook algorithm is furthermore included to improve comfort. A clear improvement of vertical acceleration can be observed
with the system installed whereas power consumption is 50-65 % less than that of a hydraulic
system.

Figure 1.4: Bose Corp. electromagnetic active suspension sytem.

The research group from Bose Corp. recognized the high power demand and low bandwidth
limitations of the hydraulic suspension systems and developed an electro magnetic suspension
system, as shown in Figure 1.4. Linear electric motors are used, making it possible to achieve a
high bandwidth to counteract the effect of road disturbances on the vehicle body [9]. According
to the manufacturer, the linear motor is also capable of delivering enough force to counteract
roll and pitch during severe cornering and braking maneuvers [10]. Due to the torsional spring
to support the vehicle weight and the possibility to regenerate energy, a power consumption (13

1.5 kW) of only one-third of the power of a cars air conditioner for the full system is claimed by
the manufacturer. However, verification of these claims has been impossible to date, since no
design details have been released nor has any commercial test been executed.
To car body

Permanent magnet array

To wheel hub
Coil spring
Three phase winding

Slotted stator
Figure 1.5: Tubular permanent magnet electromagnetic actuator in parallel with a passive spring.

Considering the low bandwidth and high power demands of hydraulic suspension and limited
performance of semi-active suspensions a novel suspension strut has been developed [11]. It
consists of a tubular permanent magnet actuator in parallel with a passive spring to support the
vehicle mass as is shown in Figure 1.5. The tubular structure gives it the capability of delivering
large direct drive forces in a small volume. Furthermore, its bandwidth is in the order of hundreds
of hertz, which is larger than required to improve comfort and handling. As a safety feature,
aluminum rings are installed in the stator. These rings provide fail-safe damping by means
of Eddy current damping. Power consumption is lower than that of a hydraulic system since
no continuous pressurization is required. Energy can even be recuperated, depending on the
amount of fail-safe passive damping and controller design [12].
All the favorable properties of the novel tubular actuator give rise to very different control
possibilities and challenges compared to currently available systems. For example the bandwidth
is great for improving comfort and handling, but might also cause resonances that other actuators
are not capable of exiting. These favorable features and potential problems make the control of
this actuator an interesting topic of this thesis.

1.1 Problem statement and objectives


Given the limitations in the aforementioned active suspension systems a tubular direct drive
electromagnetic active front suspension for a BMW 530i is developed. In this research, the control
of this active suspension will be considered. The goal is to improve the comfort or handling of
the vehicle by means of proper control of the active suspension. The main research question can
therefore be formulated as:
What performance gains in comfort and handling can be achieved with a high bandwidth electromagnetic active suspension given the constraints of maximum actuator force and suspension travel?
To answer this question, a controller for this active suspension system has to be developed. Furthermore, several research objectives have to be fulfilled to answer this question:
Given the parameters of the suspension system and the BMW, a simulation model has to
be created that represents the system correctly.
The comfort and handling performance criteria have to be defined to objectively measure
improvements.
A suitable control structure has to be developed such that, the best performance is achieved.
Real life measurements are better proof than computer simulations. A setup, will be developed that facilitates this.
Using this test setup, measurements have to be performed and will be compared with
simulations.

1.2 Literature review


Ever since the 1950s manufactures have been looking at active suspension systems such as Citrons hydropneumatic suspension. Control of these active suspension systems was picked up
by authors from the 1970s onwards [13, 14]. Since then numerous papers have been written on
the control of active suspension systems. This section aims at giving an overview of the control
topologies and results achieved with active suspensions.
In Michelberg et al. [15], a quarter car model is used to show the disadvantage of LQG control.
The author states that if the parameters of the system differ from the nominal parameters, the
controlled system can behave worse than the original plant. To counteract this the author uses
state-feedback H control in the presence of structured (parametric) uncertainties to find the
robust LQG controller. The paper shows that a trade off has to be made between the robustness
and performance requirements. To find a suitable controller one or two parameter iterations are
necessary.
De Jager [16] makes a comparison between various Linear Quadratic (LQ) controllers and
H control, based on -iterations, using a quarter car model. Weighting filters are included to
account for human sensitivity to vertical acceleration. The author finds that H control is not
suitable for the design of an active suspension controller using simple weighting filters. The bad

performance is mainly caused by the invariant point i = kt /m u which severely limits the loop
shaping possibilities the author argues. The Linear Quadratic Regulator (LQR) controller is found
5

to be more robust over Linear Quadratic Output Feedback (LGOF) or Linear Quadratic Gaussian
(LQG), the LQR controller was, however, not found practical due to the large amount of sensors
required.
Yamashita et al. [17] applies robust control in order to reduce vertical acceleration and improve handling for a full car, 7 degrees of freedom (DOF) model. A multiplicative uncertainty is
introduced to account for changing body mass and actuator uncertainty. It is assumed that the
full state is observable in the design of the controller. However, states are estimated by means of
integration and algebraic calculations in the vehicle. Experimental validation is done using a four
post shaker where all four wheels can be excited individually. A comparison is made between
a nominal performance controller and a robust controller, showing that, under the influence of
perturbations the nominal controller performs worse than the passive case. The robust controller
performs similar in the perturbed case as in the nominal case, in both cases the performance is
better than the passive vehicle. Real driving tests showed similar results as the simulations.
Fuzzy control has been applied by Sharkawy [18]. Its performance is better than standard LQR
control, however finding the output surface of the fuzzy controller is achieved by trial and error.
To overcome this, an active fuzzy controller was developed based on Lyapunov direct method
resulting in fast convergence of the parameter vector. The author, however, never discusses the
stability of this method.
Hrovat [19] discusses various models for controller synthesis. From the one DOF model using
an LQ controller he concludes that the control topology, using the suspension travel and vertical
vehicle speed as state variables, functions as a "skyhook damper". The author argues that a real
"skyhook" is physically not possible, an active device replacing the "skyhook damper" is therefore
a suitable replacement. Better performance is achieved when the active device emulates a damper
connected to a smooth inertial ground, thereby facilitating larger damping rates, but preventing
the transmissibility of road vibrations normally associated with large damping values. Using a
two DOF model, Hrovat points out that when limiting tire deflection to the tire deflection of
a passively suspended vehicle, an improvement of 11 % can be achieved. He argues that this
little improvement might not even be felt by most drivers, however, using the full potential of
an active suspension, in which adaptive tuning is used, an improvement of 67 % in comfort can
be reached. Depending on the driving conditions, the constraints on either comfort or handling
can be loosened as a function of for instance the lateral acceleration or steering angle. Hrovat
continues with a 2D model, with which he evaluates the effectiveness of preview. He finds that a
preview time of one second can reduce sprung accelerations by 50 % to 70 %. Since this time is
too long for practical implementation, the author considers a 50 ms preview, which already results
in a reduction of 30 % in tire deflection. Hrovat also discusses the stability of LQG controllers, his
conclusion, which corresponds to Doyles conclusion [20], is that no guaranteed margins exist.
In one example the LQG margins are only 0.2 d B and 18 whereas these were d B and 100
for the LQ case.
Venhovens [6] uses optimal control technology to enhance the comfort of a vehicle. First,
he uses a quarter car model to synthesize a controller. Both full and limited state feedback are
tested. Venhovens argues that limited state feedback is favorable due to the lower amount of
states that have to be determined. Furthermore, an integrator to eliminate the steady state error
that occurs with full state feedback can be avoided. The duality of emphasizing both road holding
and comfort is discussed, the author finds that not much gain can be achieved compared to a well
tuned passive system, however, he notes that the benefit of an active system is the adaptability.
Due to the high power demands of full active suspensions, Venhovens also investigates a semi
active suspension system, in which the damping value can be changed. This results in large tire
6

deflections when "skyhook" damping is used. Even with a large range of possible damping values,
no improvement in tire deflection can be achieved compared to the passive system. Venhovens
has also considered adaptive control of the suspension system, in which a predefined limit on
the tire deflection is used to tune the controller in the direction of handling or comfort. Full
car simulations showed a good correlation with the quarter car model. A benefit of the full car
model is the possible use of preview for the rear wheels, especially improving the dynamic tire
compression and suspension travel of the rear wheels. Exchange of state information between
the four corners does not improve performance noticeably.
Experimental verification of a robust controller in combination with an hydraulic actuator is
done by Lauwerys [21]. The actuator consists of a hydraulic cylinder in which continually variable
valves are places to control the flow. Furthermore, a pump is added capable of delivering a force.
RMS power consumption of this system is approximately 500 W and a delay of 6 ms is considered due to the hydraulic tubing and electrodynamic valves [22]. The dynamics of the quarter
car test setup are determined using a frequency domain approach in which the parameters and
uncertainties of the model are estimated from measurements. For this, integrated white noise
is used that represents the road disturbance. The measured outputs are the accelerations of the
sprung and unsprung mass. The nominal model is a linear approximation of the quarter car test
rig, whereas the uncertainties caused by sensor noise, non-linearities and unmodeled dynamics
are of the multiplicative type. The controller design is aimed at reducing the body vibrations
at 1.5 Hz without amplifying the body acceleration or tire force in other frequency regions. A
bandpass filter is used in which the cut-off frequencies can be chosen to reach the desired performance. The performance gain in body acceleration was found to be 50 % without serious
drawbacks in dynamic tire load variations.
Another experimental test setup was built by Lee [23]. A tubular brushless permanent magnet
motor with a peak force of 29.6 N was developed and installed on a scaled quarter car test setup.
The sprung mass of the test setup was scaled down by a factor 150 (up to 2.299 kg), whereas the
unsprung mass was scaled down by a factor 20 (to 2.278 kg). The road excitations are provided
by a cam, thereby fixing the shape of the road profile. Both the sprung and unsprung acceleration were measured as well as the suspension travel. Using these measurements, three types of
controllers were tested: A lead-lag, LQ and fuzzy controller. The LQ controller did not perform as
well as the other two controllers due to the errors in the estimated state. However, performance
gains of up to 64 % were still achieved compared to the system not in operation. The fuzzy and
lead-lag controller achieved performance gains of 77 % and 73 % respectively requiring equal
RMS currents. For both the fuzzy and lead-lag controller, the suspension travel sensor was not
used.
Many authors have considered LQ control in theory. This thesis will discuss the performance
of the LQ control topology including measurements performed on a full size quarter car test
setup. Furthermore, since a car always has variations in its parameters, this thesis will consider
robust control. The actuator is a novel tubular electromagnetic suspension system, that offers
great benefits over hydraulic systems due to its efficiency, high force density and bandwidth.
Since only the front suspension is replaced by the active suspension and measuring the road is
deemed unsuitable due to its high sensitivity for errors, no preview will be used.
7

1.3 Outline
In Chapter 2 the properties of the BMW will be introduced. After this, the suspension actuator
and sensors are discussed in detail. Having defined the actuator and BMW, a quarter car model
is created that can be used for simulations and controller development. Comfort and handling
performance criteria are defined using the quarter car model.
Having defined the performance criteria for comfort and handling, Chapter 3 defines the
objectives for control. The limitations defined by the equations of motion are shown after this.
The first control topology discussed in this chapter is linear quadratic control which finds an
optimal controller given a linearized plant. To account for uncertainties not incorporated in the
linear model, robust control is discussed.
The results of these two control topologies are determined and discussed in Chapter 4. Furthermore, the performance of the BMW is shown. In Chapter 5 the test setup is introduced to
verify the results obtained in simulations. All sensors fitted to the test setup are discussed, as well
as control of the industrial actuator that facilitates the road disturbances. A solution to a sensor
measurement error is also presented.
Using this test setup, the results of the linear quadratic and robust controller are discussed in
Chapter 6. Finally, conclusions and recommendations are drawn in Chapter 7.

Chapter 2

Actuator and car model


In this chapter the model of the actuator and car will be discussed. First the base car, a BMW 530i,
will be introduced. After this, the retrofit active suspension system will be discussed. This suspension system is equipped with three sensors used for control of the active suspension, why
these sensors were chosen and what their properties are is shown in the next section. After
this the simplified car model that will be used for controller synthesis is discussed. Finally, the
mathematical description of the road disturbance to this model is given.

2.1 BMW 530i


The BMW 5-series, see Figure 2.1, is a German built executive saloon car well known for its sportiness, agility and comfort [24]. It is available with a range of engines, from a two liter four cylinder
up top a five liter V10. In this report, a BMW 530i will be considered, which has a three liter inline six cylinder engine. With its aluminum engine, bonnet and front quarter panels it achieves a
near to perfect 50.9/49.1 front to rear weight distribution [25]. The front suspension, which will
be replaced by the active suspension, is a MacPherson strut. This system uses two suspension
arms that are connected to the bottom part of the hub and provide lateral and longitudinal fixation of the wheel. The top of the hub is attached to a suspension strut which consists of a spring
and damper in parallel as Figure 2.2 shows. Furthermore, a rebound spring is fitted inside the
damper, lowering the spring stiffness in rebound.

Figure 2.1: BMW 5 series.

Figure 2.2: MacPherson suspension system.

Table 2.1: Technical data of the BMW 530i.

Parameter
Unloaded vehicle mass
Maximum vehicle mass
Unsprung mass front (left+right)
Unsprung mass rear (left+right)
Spring stiffness
Tire vertical stiffness min-max
Weight distribution front-rear
Maximum compression (bump)
Maximum extension (rebound)

Value
1546
2065
96.6
89.8
30.01e3
3.1e5 3.7e5
50.9 49.1
0.06
0.08

Unit
[kg]
[kg]
[kg]
[kg]
[N /m]
[N /m]
[%]
[m]
[m]

Figure 2.3 shows measurements performed on the suspension strut by Janssen [25]. The
effect of the rebound spring is clearly visible from the change in gradient at -0.015 m stroke. The
damper force is clearly asymmetrical in the bump and rebound region. In compression as little
damping as possible is desired, such that the vehicle is capable of absorbing bumps. Kinematic
limitations, however, require a certain amount of damping, thereby limiting suspension travel.
Generally more rebound damping is applied to prevent abruptness in the suspension [26]. This
means that the motion of the wheel stops suddenly, thereby increasing the jerk (derivative of
acceleration) on the vehicle body.
Tires are generally considered to be non-linear both in vertical as well as cornering stiffness.
Vertical stiffness measurements have been performed on the Dunlop SP Sport 225/50R17 94W
tires on a flat plank tire tester [27], see Appendix A. This showed that, given nominal operating
conditions, the tire stiffness varies between 3.1e5 and 3.7e5 N /m. This, together with the other
car parameters is summarized in Table 2.1.
10

3000

5000

2000
4000

1000

Bump
3000

Force [N ]

Force [N ]

0
1000

Rebound

2000

Bump

2000
1000

3000
0

4000

6000
0.1

Rebound

1000

5000

0.05

0.05

2000
2

0.1

Speed [m/s]

Stroke [m]
(a) Spring Force.

(b) Damper Force.

Figure 2.3: Spring and damper characteristics of the BMW 530i front suspension.

2.2 Active suspension system


2.2.1 Actuator
To generate a force for suspension control an electro-magnetic actuator has been designed [28,
29]. The actuator has been designed such that it is a retrofit for a BMW 530i McPherson front
suspension strut. Performance specifications have been derived from measurements performed
on the Nrburgring in Germany. There it was found that a peak and RMS force of respectively
4000 N and 2000 N was necessary to eliminate the vehicle roll angle. The author also commented that these driving conditions are not very common. A duty cycle of 50% is therefore
proposed, resulting in an RMS force of 1000 N.
The electro-magnetic actuator is a tubular slotted three-phase permanent magnet actuator [29].
A graphical representation is given in Figure 2.4 with a detailed view in Figure 2.5. It can be seen
that a quasi-Halbach array has been chosen. This topology offers the highest force density [30]
by focussing the magnetic field into the actuator. An external magnet array has been chosen to
increase magnetic loading. Furthermore, copper losses are reduced due to the smaller circumference of the coils. Another benefit is the absence of moving wires since the power electronics
are situated on the sprung mass. In this stator, angular coils are fitted, such that, according to
Lorentz
Z
Z
E V Fz =
JE Bd
J Bv d V,
(2.1)
FE =
an axial force is generated. The aluminum rings are fitted such that Eddy currents are induced
when the actuator moves. This provides passive damping since the Eddy current creates a force
opposing the original movement. The spring that is placed in parallel with the actuator compensates for the mass of the car such that no continuous power is required to levitate the car. Fur11

Sprung acceleration sensor


To car body

Aluminium ring

Coil spring

Halbach magnet array

Sliding bearing

Ls

Laser Sensor

Unsprung acceleration sensor

Ri

Three phase winding

Rm

Bump stop
Rr
Ro

Figure 2.4: Electro-magnetic actuator cross section.

thermore, the stroke of the actuator is chosen such that is equal that of the passive suspension
strut. The linear guidance of the stator is done by means of a linear sliding bearing that is fitted
over the entire length of the magnet array. Relevant parameters are summarized in Table 2.2.
Figure 2.6 shows the electric motor model composed of a voltage source (Vi ), resistor (Ri ), inductor (L i ) and back-EMF (E i ). The differential equation that describes this model is formulated
as
d Ii
(2.2)
Vi = E i + Ri Ii + L i
dt
where the subscript i denotes phase a, b or c, furthermore, Ri denotes the resistance per phase
and L i is the inductance. The current in these phases is given by


z

Ia = i sin
+
(2.3)
p


z 2
Ib = i sin

+
(2.4)
p
3


z 4

Ic = i sin

+
(2.5)
p
3
12

z
Flux lines

Aluminium rings

Hallbach magnet array


Figure 2.5: Electro-magnetic actuator detailed view of three phases.

Table 2.2: Actuator parameters

Parameter
Rs
Rm
Rr
Ro
Ls
p
Lb
Lrb
ks
FR M S
m trans
m stat

Value
26.925
28
36
39
400.4
7.7
60
80
30.01
1000
7
8

Unit
[mm]
[mm]
[mm]
[mm]
[mm]
[mm]
[mm]
[mm]
[N /mm]
[N ]
[kg]
[kg]

Description
Stator radius
Inner magnet radius
Outer magnet radius
Outer translator radius
Stator length
Pole pitch
Bound stroke
Rebound stroke
Spring stiffness
Maximum RMS actuator force
Actuator translator mass
Actuator stator mass

13

Where, z = z s z u is the displacement of the actuator (suspension travel), p the pole pitch, i the
amplitude of the current and speed dependent commutation. The EMF E i is given by
E i = k Ei v

(2.6)

with k Ei the EMF gain and v = vs vu the speed of the actuator. Assuming that i and v are
independent, the force delivered by the actuator is given by:
Fact = Fcurr ent + Fdamp = k I i + dv

(2.7)

with k I the force gain and d the damping coefficient. The measured force as a function of current
can be seen in Figure 2.7. Also visible is a linear approximation of this measured force

Fact = ki i = 115i.

(2.8)

This approximation is valid up to 2000 N which means ki can be used in simulations. The
amplifier, using a 2 k H z current control loop, makes sure that this force is really generated.
Figure 2.8 shows the contribution fail-safe electromagnetic damping. Due to final inductance
of the rings, the damping has a regressive character. Also visible is an approximation of this
non-linear damping, this can can be formulated as
Fdamp = 1341 arctan (0.985v) .

(2.9)

Using simulations, the occurrence of a certain damping value is tested when driving over a certain road. As Figure 2.9 shows, large damping values occur much more than small values, this
corresponds to relatively small suspension speeds, v. The average damping value is 1450 N s/m.

Ri

Li

Vi

Ei

Figure 2.6: Model of an electric motor.

14

3500
Measured
3000

F = 115 i

Fact [N ]

2500

2000

1500

1000

500

10

15

20

25

30

Current [A]
Figure 2.7: Actuator force vs current, measured and approximated with a K i of 115 N /A.

1400
Measured
1341arctan(0.985v)

1200

Fdamp [N ]

1000

800

600

400

200

0.5

1.5

Speed [m/s]
Figure 2.8: Eddy current damping, measured and approximated with Fdamp = 1341 arctan (0.985v).

15

1
0.9

Occurrence []

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

Mean ds
1000 1100 1200 1300 1400 1500 1600

ds [N s/m]
Figure 2.9: Damping occurrence on rough road.

2.2.2 Sensors
As Figure 2.4 shows, three sensors are fitted to the actuator. This set of sensors is most commonly
used in literature [31]. The sprung acceleration sensor measures the acceleration of the vehicle
body, this sensor is fitted, because the acceleration of the vehicle body is a direct measure for
comfort. The second sensor fitted is a laser sensor that is used to measure the suspension travel.
Since the suspension travel is directly coupled to the commutation of the actuator, it is important
to measure this value directly. Moreover, since suspension travel is limited, this sensor gives a
good indication of the state of compression. Appropriate action can be taken if the system gets
close to its limits. The third sensor fitted is a 50 g acceleration sensor, this is used to measure the
acceleration of the unsprung mass. Since it is impossible to measure the absolute compression of
a tire due to the unpredictable nature of a road surface, this measurement is the most convenient
to determine the state of the tire. This unpredictable nature is best expressed by the example of
driving over a brick or a carton of milk. Both appear to be solid for a sensor, whereas the carton
of milk will be compressed easily by the tire, the brick might cause damage.
Obviously, measurements are noisy. Each sensor has a certain noise level, based on measurements and manufacturer specifications, the noise levels have been determined as Table 2.3
shows. Implementation in simulations is done by multiplying a white noise signal by a gain, Wni ,
to achieve the sensor noise as Figure 2.10 shows.
As Van de Wal [31] indicated, good control is also possible with only the acceleration sensors.
In theory, the suspension travel could be estimated through integration, however, due to possible
sensor drift and the high importance of correct commutation the suspension travel is measured
directly.
16

Table 2.3: Sensor noise.

Sensor
Sprung acceleration sensor
Sprung acceleration sensor
Suspension travel sensor

White noise

Wni

Deviation
0.024 m/s 2 RMS
0.178 m/s 2 RMS
0.002 m RMS

Sensor noise

Figure 2.10: Implementation of sensor noise.

2.3 Simplified car model


Having defined the baseline car and tubular actuator, a model has to be constructed that represents these parts. For this, the actuator is installed in the BMW as Figure 2.11 shows. Only one
quarter of the car will be used as simulation model, this will be explained in more detail in the
next section. Section 2.3.2 will discuss the road input to the model.

Figure 2.11: Actuator installed in BMW.

2.3.1 Quarter car model


Models to asses the vertical dynamics of a vehicle exist at various different levels. From very
simple, 1 DOF models [19] up to non linear large number of DOF models [32]. It has, however,
been shown [33] that a 2 DOF car model represents the vertical dynamics of a vehicle accurately
enough to predict the comfort and tire compression of the vehicle.
A quarter car model represents one corner of the vehicle for which only the vertical dynamics
are considered. Figure 2.12 shows a graphical representation of the quarter car including an
actuator. Here, m s is the sprung mass of the vehicle, m u is the unsprung mass, this is usually
made up out of the weight of the rim, tire, brake and part of the suspension. The stiffnesses and
17

damping are denoted by ks , kt and ds respectively, with kt the vertical tire stiffness. The degrees
of freedom are the displacement of the sprung (z s ) and unsprung mass (z u ). The displacement of
the road zr is prescribed by the road profile as is discussed in more detail in section 2.3.2. Finally,
the actuator force is denoted by Fact . The equations of motion are given by:
m s z s = ks (z s z u ) ds (z s z u ) + Fact

(2.10)

m s z u = ks (z s z u ) + ds (z s z u ) kt (z u zr ) Fact

(2.11)

The vertical acceleration (zs ) is a good indication of the ride comfort of a car, humans, however,
are only sensitive to vibrations up to a certain frequency. To take this into account, a weighting
filter according to ISO2631-1 will be used [34]. Figure 2.13 shows the frequency dependent weighting function. As can be seen, humans are most sensitive to vibrations in the 4-10 H z range, with
fast decreasing sensitivity beyond this range. At lower frequencies humans are also less sensitive,
however, motion sickness occurs at roughly 0.125 H z for humans that are sensitive for this.

zs, zs, zs
ms

ks

Fact z, v
zu, zu, zu

ds

mu
kt

zt
zr

Figure 2.12: Quarter car model.

Figure 2.12 also shows the tire compression (z t ), this is a good measure for handling, since
side force can be maximized when the vertical force changes remain minimal. This is due to
relaxation effects being minimized when dynamic tire compression is minimized.
The third important performance parameter indicated in Figure 2.12 is the suspension travel (z).
This is defined by the available space in the suspension system, for the BMW this is 0.06 m in
compression and 0.08 m in extension. One can imagine that if more space is available, a more
comfortable car can be built, since the distance before the suspension hits the bump stops is
larger en therefore less damping or a lower stiffness can be chosen. This parameter is therefore
of high importance to ensure a fair comparison between the passive and active suspension.

2.3.2 Road input


A vehicle is subjected to a lot of disturbances while driving. Typically, two types of disturbances
can be identified: stochastic irregularities and deterministic disturbances. Stochastic irregular18

1.1

Magnitude ISO2631-1 []

1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
1
10

10

10

10

Frequency [Hz]
Figure 2.13: Human body sensitivity to vertical vibrations, ISO2631-1.

ities describe normal driving conditions, from speedbumps to random vibrations reflecting the
surface quality of the road. Measurements [35] have shown that these stochastic irregularities can
be represented accurately by colored noise resulting from the application of a first order filter to
a white noise signal w [6].
1
zr + zr = w
Vx

(2.12)

Here zr is the vertical road input, Vx is the forward speed of the vehicle. The parameter defines
the cut-off frequency and thereby the shape of the road irregularities. Figure 2.15 shows measurements performed on a smooth and a rough road. Table 2.4 shows the parameters that have been
used to create the simulated road profiles. Here, the first order low-pass output is multiplied with
to achieve the correct road amplitude. This gain is based on a sample time (ts ) of 0.001 s in
Matlab-Simulink and has to be changed when this sample time is changed.
Table 2.4: Typical road parameters.

Road type

[rad/m]

Vx [m/s]

Smooth

0.2

30

Rough

0.8

7.5

19

q
[]
0.05

0.001

q ts
0.125 0.001
ts

0.05

zr [m]

0.04
0.03
0.02
0.01
0
0.5

1.5

x [m]
Figure 2.14: Profile of the speed bump used as deterministic disturbance.

10

10

PSD zr [m2 /Hz]

10

Ve
ry

10

Av

Ve

ry
er
P
Go age Poo oo
r
r
od

od

10

Measured smooth road


Measured very rough pavement
Simulated smooth road
Simulated rough road
ISO8608 classifications

10

10

Go

10

10

10

Spatial frequency [1/m]

Figure 2.15: Power spectral densities of different road types.

20

10

Figure 2.15 shows this first order low-pas filter including the parameters compared to the
measurements in spatial frequency. This spatial frequency is determined as:
ns =

f
Vx ,

(2.13)

with f the frequency and Vx the forward velocity. As is visible, the measured smooth road is
generally classified as a very good road according to the ISO8608 [36] with a large peak at small
spatial frequencies, most likely caused by a very low frequency wave in the road surface. The
simulated smooth road has been chosen such that it is also classified as a very good road and that
it matches the measured smooth road in at higher spatial frequencies. For the simulated rough
road, the maximum capabilities of the test setup are considered, as will be discussed in Chapter 5,
this results in a road that is classified as a good to average road by the ISO8608 criterion.
As deterministic disturbance a 30 mm high speed bump is chosen as Figure 2.14 shows. It has
a 45 degree angle relative to the road surface. Tire enveloping behavior over this road disturbance
will not be taken into account for simplicity reasons.

2.4 Summary
In this chapter the BMW 530i is introduced as the base vehicle. Its spring and damping characteristics are found to be non-linear. This is also true for its tire properties, which have been measured
on a flat plank tire tester. The active suspension that will be used consists of a tubular electromagnetic actuator in parallel with a passive spring to support the vehicle weight. Aluminum rings are
fitted to provide passive fail-safe Eddy current damping.
To asses improvements in comfort and handling a quarter car vehicle model is used. For the
comfort, vertical acceleration of the sprung mass will be used. To account for human sensitivity
the ISO2631-1 criterion is used. The quality of handling is expressed by dynamic tire compression. An important constraint that limits the amount of comfort that can be achieved is the
suspension travel. As input to the model both random and deterministic disturbances will be
used. The random road is represented by white noise filter by a first order low-pass filter. As
deterministic input a three centimeter high speed bump is used.

21

Chapter 3

Control of the active suspension


Given the high bandwidth and low power consumption of the active suspension introduced in the
previous chapter a controller has to be developed that makes uses of these favorable properties.
This chapter will deal with the design of the controller, based on the quarter car model with a
random road as input. First the control objectives will be formulated. After this limitations for
control of the active suspension will be given. Finally, two different control approaches will be
explained, being Linear Quadratic control (LQ) and Robust Control (RC).

3.1 Control objectives


For both control topologies, the same objectives and constraints hold. The ride comfort objective
is defined by the RMS vertical acceleration zs weighted by the ISO2631 criterion. The smaller this
value the better.
Ride comfort = RMS (W I S O2631 z s ) = RMS (z sw )
(3.1)
Here, the ISO2631 weighting filter is approximated by a fifth order transfer function as suggested
by Zuo [37]
87.72s 4 + 1138s 3 + 11336s 2 + 5453s + 5509
W I S O2631 = 5
.
(3.2)
s + 92.69s 4 + 2550s 3 + 25969s 2 + 81057s + 79783
The handling objective is defined by the RMS dynamic tire compression z t , minimization of this
variable maximizes the lateral and longitudinal forces.
Handling = RMS (z t ) = RMS (z u zr )

(3.3)

Furthermore, suspension travel z is limited due to constraints of the suspension. The allowed
suspension travel of the active system is limited to the maximum value of the passive BMW
suspension to make a fair comparison.
max (z Active ) max (z B M W )

(3.4)

Finally, as was indicated in Section 2.2, the actuator has been designed for a maximum RMS force
of 1000 N . So when designing the controller the RMS value should stay below this value.
RMS (Fact ) 1000 N
22

(3.5)

3.2 Duality of control objectives


Independent of the chosen control topology, constraints exist that limit the performance of any
active suspension [38]. To make this clear, consider the equations of motion of the quarter car,
see (2.10) and (2.11). Converting them into Laplace domain and summing them results in
m s z s s 2 + m u z u s 2 = kt (z u zr ) ,

(3.6)

in which no suspension forces can be seen, the consequences of this will be discussed later on.
The transfer functions between the input zr and outputs z s , z u zr and z s z u are defined as
Hzs/zr = s 2 Hzs/zr =

kt (ds s + ks ) s 2


.
m s s 2 + ds s + ks m u s 2 + kt + m s s 2 (ds s + ks )

(3.7)

Furthermore,
H(zuzr )/zr


m u s 2 m s s 2 + ds s + ks + m s s 2 (ds s + ks )


=
m s s 2 + ds s + ks m u s 2 + kt + m s s 2 (ds s + ks )

(3.8)

kt m s s 2

.
m u s 2 + kt + m s s 2 (ds s + ks )

(3.9)

H(zszu)/zr =

m s s 2 + ds s + ks

Now (3.6) can be rewritten as



m s s 2 Hzs/zr + m u s 2 + kt Hzu/zr = kt .

(3.10)

If s = j, it can be seen that if equals


s
W H =

kt
,
mu

(3.11)

this reduces to

kt
Hzs/zr = kt .
(3.12)
mu
This is frequency is graphically illustrated in Figure 3.1 where the transfer functions of a controlled and passive quarter car model can be seen. The amplitude of Hzs/zr can now be derived as
being
mu
Hzs/zr |s= jW H =
(3.13)
ms
and the amplitude of
kt
(3.14)
Hzs/zr |s= jW H = .
ms
This frequency (3.11) is called the wheel-hop frequency [39]. Rewriting (3.6) relates the suspension
travel to the sprung motion

(m u + m s ) s 2 + kt Hzs/zr kt
H(zszu)/zr =
.
(3.15)
m u s 2 + kt
m s

From this equation it follows that

s
RS =

kt
.
ms + mu
23

(3.16)

The amplitude of (3.15) becomes


H(zszu)/zr |s= j RS =

mu + ms
.
mu

(3.17)

This is called the rattle space frequency [39] and can not be influenced by the suspension force
as (3.6) showed. Again, this frequency is illustrated in Figure 3.1. Equation (3.6) can also be

Sprung acceleration

10

Hzs/zr [dB]

10

W H

10

10

H(zszu)/zr [dB]

10

10
Suspension travel

10

10

10

H(zuzr)/zr [dB]

10

10
Dynamic tire load

15

10

10

10

Passive
Controlled

10

10

10

Frequency [Hz]
Figure 3.1: Invariant points in sprung acceleration and suspension travel.

rewritten into a transfer function that relates tire deflection and the body motion transfer function
H(zuzr )/zr


s 2 m u + m s Hzs/zr
=
.
m u s 2 + kt

(3.18)

There is no point in frequency that can not be influenced by an active suspension, except at
= 0, where H(zuzr )/zr = 0. These invariant points thus show that, independent of the control
approach, certain points can not be influenced.
Equations (3.10), (3.15) and (3.18) furthermore show that if one were to optimize one performance variable, concessions have to be made in terms of the other variables. Hedrick [38] points
out by considering the derivative of (3.6) that, although the transfer functions are coupled, at low
24

frequencies an improvement in both ride comfort and tire deflection can be made. At high frequencies, the author found that when requiring an improvement in tire deflection, the increase
in vertical acceleration is large. The second point that Hedrick discusses, is that in general, for
lower vertical acceleration, a large increase at low frequencies and near the wheel hop frequency
can be observed. A physical explanation for this phenomenon is that an improvement in vertical
acceleration requires a low spring stiffness, which results in a large suspension travel.

3.3 Linear quadratic control


A time invariant linear system can always be stabilized by a linear feedback if it is fully controllable [40]. By choosing the poles far in the left half of the complex plane, infinitely fast convergence can be achieved. To make this possible, large control amplitudes are necessary. Since this
is physically impossible, there exists a limit on how far the poles can be moved to the left. To find
an input that suffices both the requirement of fast control and does not require infinite control
power, an optimization problem has to be solved. A very useful criterion is the quadratic integral
criterion
Z
t1

J = lim

t1

t0

ycT (t) Qyc (t) dt

(3.19)

with yc the controlled variable and Q a diagonal non-negative weighting matrix containing the
weighting factors. Given the objectives of dynamic tire compression z t , sprung acceleration zs
and suspension travel z the output, yc , can be formulated

z u zr
zt
= z s = C x + Du.
z s
yc =
zs zu
z
With x the state vector
x=

zs

z s

zu

z u

zr

T

(3.20)

(3.21)

and u the input Fact , C and D can be written as:

0
0

ks
ds
C =
m m
s
s
1
0

1
ks
ms
1

0
ds
ms
0

0
1

D=
m .
s
0

(3.22)

(3.23)

Now consider (3.20), substituting this into (3.19) results in


Z
J = lim
Z

t1

t1 

t0

xT

uT

t1

(C x + Du)T Q (C x + Du) dt =



 C T QC C T Q D 
x u dt
T
T
D QC D Q D
t0

25

(3.24)

this is abbreviated as
Z
J = lim

t1

t1 

t0

Qc
NcT

Nc
Rc

dt.

(3.25)

Calculus of variations leads to the state feedback


u (t) = K x

(3.26)

with K
K = Rc1 NcT + B T P

(3.27)

and P (P = P T > 0) the solution of the Riccati equation


P(A B Rc1 NcT ) + (A B Rc1 NcT )T P P B Rc1 B T P + Q c Nc Rc1 NcT = 0,

(3.28)

with A and B being the state matrices defined by (2.10), (2.11) and (2.12) as

A=

0
ks

ms
0
ks
mu
0

1
ds

ms
0
ds
mu
0


B=

0
0
0
ks
ds
0
ms
ms
0
1
0
(ks + kt )
ds
kt

mu
mu mu
0
0
av
1
ms

1
0
ms

(3.29)

T
0

(3.30)

The foregoing analysis has omitted the fact that a white noise disturbance is present as road input.
Consider the state equation
x = (A B K ) x + w
(3.31)
with w the white noise disturbance. It can then be proven [40] that the solution does not alter,
except to increase the minimum value of (3.19).
A necessary requirement is that the system is controllable. Controllability can be checked by
determining the rank of


R = B AB A2 B . . . An1 B
(3.32)
if this rank is equal to n, with n the size of the state, the system is controllable. The system
defined here is not fully controllable, which is expected, since the actuator cannot influence the
road displacement zr . A suitable controller will, however, still follow from the above Riccati
equation as the system is still stabilizable.
Since LQ control only considers a linear model, constant parameters have to be chosen. Table 3.1 shows the parameters selected. The sprung mass, m s , is based on the empty weight of the
car plus two passengers and half a tank of fuel. It is assumed that the weight of the passengers
and fuel is evenly distributed over the car. For the damping value, ds , the average value as was determined in Section 2.3.1. Finally, for the tire stiffness, kt , the total load of m s + m u is considered.
From this the stiffness is derived from Figure A.2.
26

Table 3.1: Car parameters considered for LQ control.

Parameter
ms
mu
ks
ds
kt

Value
395
48.9 + m trans
30.01e3
1450
3.37e5

Unit
[kg]
[kg]
[N /m]
[N s/m]
[N /m]

Description
Sprung mass
Unsprung mass + actuator mass
Spring stiffness
Damping
Tire stiffness

Given the three output variables, a weighting matrix Q can be defined as

q1 0 0
Q = 0 q2 0
0 0 q3

(3.33)

with q1 , q2 and q3 emphasizing tire deflection, sprung acceleration and suspension travel respectively. These weighting factors are solved by means of a constrained nonlinear optimization
algorithm, fmincon. The objective function minimized is
R M S (z t )
R M S (z s )
 + (1 )

O = 0.5
R M S z sp
R M S zt p

!
(3.34)

where comfort ( ) or handling (1 ) is emphasized depending on the choice of . Here, z sp


is the performance of the BMW. Suspension travel and actuator force are used as constraints for
the optimization.
In this section it is assumed that the full state is measurable. On the quarter car test setup
this will not be a problem since the full state is measurable. On a real car, this will, however, be
a problem. The state will therefore have to be estimated. The resulting problem is the Linear
Quadratic Gaussian (LQG) control problem.

3.4 Robust control


As Doyle has shown [20], stability margins can not be guaranteed with LQG control. It is therefore
necessary to explore other control topologies, that can guarantee stability, even with an uncertain
plant. H -control seems to be able to guarantee this stability [41]. Using the structured singular
value, defined as
(M)1 min{ (1) | det (I M1) = 0 for structured 1}
1

(3.35)

DK-iteration can be performed to synthesize a -optimal controller. Here, M is considered the


part of the plant connected to the uncertainty matrix 1. The idea is to find a controller that
minimizes the peak value over frequency of the upper bound
(N ) min D N D 1
DD

27

(3.36)

namely,



min min |D N (K ) D 1 | .
K

DD

(3.37)

Here, K is an H -controller that is synthesized while D is kept fixed. D is a matrix that is found
by minimizing D N D 1 ( j) with N fixed. Finally, each element of D( j) is fitted to a stable
and minimum phase transfer function D(s). The matrix N is the generalized plant defined as the
lower fractional transformation of P and K , with P the plant and K the controller.
N = Fl (P, K ) P11 + P12 K (I P22 K )1 P2 1

(3.38)

The iterations continue until |D N D 1 | < 1 or the H -norm no longer decreases. The order of
the controller resulting from this process is equal to the number of states in the plant plus the
number of states in the weighting filters plus twice the number of states in D(s) [42].

3.4.1 Model
The quarter car model used to design the robust controller is similar to the model introduced in
section 2.3.1, however, various uncertain parameters and weighting filters are now included in
the model as is shown in Figure 3.2. Uncertainties in the model can have several origins [41];
There are always parameters in the linear model that are only known approximately or
are simply wrong. Furthermore, parameters can vary due to non-linearities (such as the
damping coefficient) or changes in operating conditions (such as changing tire stiffness, as
a function of load and inflation pressure, and sprung mass).
Measured signals are imperfect, sensor noise and discretization errors can cause the signal
to deviate from its real value. This can give rise to uncertainty in the input. For the three
sensors present in the test setup the noise levels are summarized in Table 3.2 together with
the parametric uncertainties.
At high frequencies the structure and model order are unknown. Therefore uncertainties
will always surpass 100 % at some frequency. Good examples of this are the chassis resonances beyond 30 H z and the natural frequencies of the tire, which typically start at 35 Hz
[43] and beyond.
A simpler model can be chosen in favor of a very complex model, the neglected dynamics
can be incorporated as uncertainties.
Controller implementation may differ from the one obtained by solving the synthesis problem. To account for controller order reduction, one may include some uncertainty.
Output uncertainty can influence the performance of the system. Particularly deviations in
the actuator introduced in Section 2.2 such as hysteresis and temperature dependency can
influence the performance of the actuator and thereby the performance of the system.
In Figure 3.2 the uncertain sprung mass, tire stiffness and damping are included in the perturbed plant as uncertain parameters that can vary within a certain range. It is furthermore assumed that beyond 30 H z the dynamics of the system are not known completely. A multiplicative
uncertainty is therefore included
Pp = P (I + WU nmod 1 I )
28

(3.39)

Controlled outputs
zt
White noise

1
s/av+1

Wo1

Weighted dynamic
tire compression

Pertubed plant, Pp

Wi1
zr

zs

Fact

Wo3
ISO2631

Weighted sprung
acceleration

zu
zs zu

Wo4
Weighted suspension
travel

Wn1
noise
Wn2

Wn3

noise

noise

Controller
inputs

Controller

Wo2
Fact

Weighted actuator
force

Figure 3.2: Model used for DK-synthesis.

Pp
wU nmod

I
Parameteric
perturbed
plant

Unmodeled dynamics

Figure 3.3: Unmodeled dynamics.

29

Table 3.2: Uncertainties of the quarter car model.

Parameter
Sprung mass
Tire stiffness
Damping coefficient
Sprung acceleration sensor
Sprung acceleration sensor
Suspension travel sensor

Type
Parametric uncertainty
Parametric uncertainty
Parametric uncertainty
Sensor noise
Sensor noise
Sensor noise

Mean value
395.3 kg
3.4e5 N /m
1450 N s/m
-

Deviation
42.77 +75.38 kg
0.3e5 N /m
550 +250 N s/m
0.024 m/s 2 RMS
0.178 m/s 2 RMS
0.002 m RMS

as Figure 3.3 shows. Here WU nmod is defined as


WU nmod =

1
s 2 + 20.707
s+1
2 30
(2 30)2
.
1
s 2 + 20.707
s+1
2 400
(2 400)2

(3.40)

Sensor noise is included in the form of additive uncertainties to the measured sprung acceleration, unsprung acceleration and suspension travel. This additive uncertainty is in the form of
white noise multiplied by a weighting function Wni , with i ranging from 1 to 3. The four weighted
and controlled outputs, dynamic tire compression, actuator force, sprung acceleration and suspension travel are used in the DK-synthesis. Inputs to the controller are the sprung acceleration,
unsprung acceleration and suspension travel. Table 3.2 summarizes the uncertainties. The choice
of the weighting filters will be discussed in more detail in section 3.4.3.

3.4.2 Robustness requirements


The main requirement of the controlled system is performance, however, stability is also of importance. This stability requirement can be divided into nominal stability (NS) and robust stability (RS). Nominal stability can be shown by determining the poles of the controlled system
with 1 = 0. This is shown in Figure 3.4 together with the pole plot of the allowed perturbations.
It can be seen that all poles are in the left half plane, which means that all perturbations of the uncontrolled system are stable. Varying sprung mass or tire stiffness results in a shift of the poles
in vertical (imaginary-axis) direction. Changing damping results in the real value of the poles
changing. A damping, ds , of zero will results in poles on the imaginary axis, however, this will
not occur in practice. Furthermore, it is assumed that 1 is stable. Robust stability means that the
controlled system is also stable for all perturbed plants. For this the N 1 structure is considered
as is shown in Figure 3.5. The transfer function from exogenous inputs u to outputs yc is defined
as
(3.41)
Fu (N , 1) = N22 + N21 1 (I N11 1)1 N21 .
With N11 the coupling of the plant to the disturbances 1 and N22 the nominal plant. Nominal
stability already proves that the whole of N must be stable, therefore the only source of instability
can be the feedback term (I N11 1)1 . Thus when the system is nominally stable, the stability
of the perturbed system is equal to the stability of the M1-structure shown in Figure 3.6 with
M = N11 . The stability of the M1-structure can be proven by applying the Nyquist criterion.
This results in the requirement that the M1-structure is stable for all allowed perturbations with
30

(1) , if and only if


(M ( j)) < 1, ,

(3.42)

with the structured singular value.


st

nd

1 pole pair

100

20

ds

0
20

Imaginary-axis []

Imaginary-axis []

60

40

4
2

Nominal plant
2
4
6

80

8
15

10

10
2.5

Perturbed plants

60

100
20

ds

m s and kt

m s and kt

80

40

pole pair

10

Real-axis []

1.5

0.5

Real-axis []

Figure 3.4: Pole plot of quarter car model and perturbations.

If all stability requirements are satisfied, the robust performance (RP) demands have to be
fulfilled. These demands indicate whether the controlled system achieves better performance
than the uncontrolled system under the influence of the uncertainties. For this the worst case
gain from exogenous inputs w to outputs z is calculated over all frequencies for the controlled
and uncontrolled plant. Robust performance is then achieved if
1 (Nc ( j))
< 1,
1 (Nu ( j))
where is calculated with respect to the matrix


= 1 0
1
,
0 1P

(3.43)

(3.44)

1 contains the true uncertainties and 1 P is a full complex matrix with the same size as the
number of outputs of P stemming from the H -norm performance specification.
Robust performance, however, is not required for all outputs. If for instance sprung acceleration is emphasized, tire deflection does not have to perform robustly. It is only required to be
stable.

31

y
N11

N12

N21

N22

yc
N

Figure 3.5: N 1-structure used for robust performance analysis.

M
Figure 3.6: M1-structure used for robust stability analysis.

32

3.4.3 Weighting filters


Weighting filters can be used to shape input signals, such as the road disturbance discussed in
section 2.3.2 or set performance goals for the output. Depending on the shape and amplitude
of performance filters, the frequency response of the outputs can be influenced. For instance
by choosing the ISO2631 criterion as a weighting filter for the sprung acceleration, frequencies
between 4 and 10 H z are emphasized much stronger than frequencies outside of this range.
Below, the individual weighting filters will be discussed in more detail.
Sprung acceleration
Humans are most sensitive for vertical vibrations between 4 and 10 H z [34]. The ISO2631-1
standard has been created to take this into account when evaluating suspension performance. For
simulation purposes, this frequency dependent weighting has to be converted into a continuous
time transfer function, this has been done by Zuo et al. [37] and is shown in Figure 3.7. As can
be seen, up to fifth-order fits have been created, however, to keep the controller order as low as
possible, it has been decided to use the second order fit which is expressed as
Wzs (s) = wzs

s2

86.51s + 546.1
,
+ 82.17s + 1892

(3.45)

with wzs a gain that determines the importance of this weighting filter. A problem that occurs with
the use of this weighting filter is that at high frequencies the filter has a very low gain, thereby
allowing the vertical acceleration to be extremely large and causing instability. To prevent this,
Wzs is multiplied by a first order PD-filter at 200 H z. The deviation from the ISO2631-1 standard
at low frequencies is not considered to be a problem, since this is not the most important region.
Figure 3.11 shows the normalized sprung acceleration weighting filter together with the other
weighting filters.
Dynamic tire load
In literature [33, 44] dynamic tire compression is often used as an indication for the quality of
the cars road holding. This is a valid assumption as vertical load influences the lateral force a
tire can develop [43]. However, due to tire relaxation effects, this quick variation of vertical tire
compression might not have such a major influence on lateral tire force. To investigate this, a tire
model proposed by Pacejka [45] will be used. In Figure 3.8 the model is shown. The contact patch
of the tire is connected to the rim via the lateral stiffness, kr . This contact patch is only allowed to
move in y-direction with respect to the rim, therefore, Vx is assumed to be equal to Vx0 . The force
Fy in the contact patch is calculated using the Magic Formula tire model

Fy = D sin C arctan B 0 E B 0 arctan B 0
(3.46)
with 0 the side slip angle of the contact patch and parameters B, C, D and E load dependent
Magic Formula parameters. These are explained in Appendix A. The mass m c and damping dr of
the tire contact patch proposed by Pacejka [45], together with the measured stiffness kr are shown
in Table 3.3. As input to the model a fixed side slip angle is chosen as well as the vertical force
variation, derived from the quarter car model. The equation of motion can now be derived as

0
m c Vsy + dr yr + kr yr = Fy 0 , Fz
(3.47)
33

ISO 2631

1.4

Value weighting function []

2nd order fit


3rd order fit
1.2

4th order fit


5th order fit

0.8

0.6

0.4

0.2
1

10

10

10

10

Frequency [Hz]
Figure 3.7: Low order continuous time fit of the ISO2631-1 vertical acceleration criterion.

Table 3.3: Relaxation model parameters.

Parameter
mc
kr
dr

Value
1
271
5400

Unit
[kg]
[k N /m]
[N s/m]

Description
Contact patch mass
Lateral tire stiffness, measured on a non-rolling tire
2% damping to prevent instability

with the force on the wheel center expressed as


Fy = kr yr + dr yr .

(3.48)

yr = Vsy0 Vsy

(3.49)

The displacement yr can be found by

finally, the side slip angle 0 can be found by solving


0

Vsy0
|Vx |

(3.50)

Depending on the side slip angle, the variation in vertical force influences the variation in
lateral force differently as is shown in Figure 3.9. In this figure, the lateral force, Fy , is considered
34

Vx

Vx0

V0
0

kr

F
Fy

dr

S0

yr
Figure 3.8: Enhanced non-linear transient tire model.

as a function of varying vertical force, from which the PSD is represented by the black line. As
can be seen, at small side slip angles, the variation in lateral force does not appear as severe in
the vertical force as it does with greater side slip angles. Figure 3.10 shows this in more detail. In
this figure, the power spectral density of the variation of the lateral force is divided by the power
spectral density of the vertical force. It again shows that at small side slip angles, the Fz variation
does not influence the lateral force at high frequencies as much as at low frequencies. All these
figures are created using a smooth road profile and a speed of 15 m/s. Different road profiles and
speeds, however, give similar results. The drop-of can be approximated by a first order filter
Hzt (s) = 0.3

1
s+1
2 12
.
1
s+1
2 0.6

(3.51)

This filter can be used when designing the weighting filters since small side slip angles will most
likely be considered when emphasizing comfort. Therefore the weighting filter will be similar
to (3.51),
1
s+1
2 12
Wzt (s) = wzt 1
(3.52)
s+1
2 0.6
allowing for more vertical displacement at higher frequencies. Here, wzt stands for a gain factor
with which one can emphasize dynamic tire compression. Figure 3.11 shows the normalized
dynamic tire compression weighting filter together with the other weighting filters.
Actuator force
The actuator has been designed such that its limits lie beyond the frequencies that are of interest
for improving comfort. Its limiting frequency is defined by the voltage equation
Vs = R I + E + L

dI
Fmax
1Fmax
=R
+ Kr v + L
f
dt
Ki
Ki
35

(3.53)

10

10

PSD Fy [N 2 /Hz]

10

10

= 1
= 2
= 3

10

= 4
= 5
= 10

= 15

10

= 20
F var
z

10

10

10

10

10

Frequency [Hz]
Figure 3.9: PSD of variation in lateral force as a function of varying vertical force for a smooth road at 15 m/s.

where a velocity of of 1 m/s and a switching of force between 1000 N and -1000 N as peak
operating conditions are assumed. The actuator parameters R, L and motor constants are taken
from Gysen et al [12]. Taking into account the 170 V voltage limit of the amplifier, this results in
170 = 1.7 5.4 + 123.3 + 0.01 10.8 f
170 9.18 123.3
f =
= 347H z.
0.01 10.8

(3.54)
(3.55)

It is, however, not necessary to have the actuator deliver force up to this frequency, since no
benefit in either comfort or handling can be achieved at such frequency. It is therefore chosen to
limit the actuator force beyond 30 H z, since tire [43] and chassis [46] resonances can be excited
beyond this frequency. The resulting weighting filter is a second order, limited high pass filter
(skewed notch):
21
1
s 2 + 2
s+1
30
(2 30)2
W Fact = w Fact
(3.56)
1
2
s 2 + 22200
s+1
(2 200)2
Here, 1 and 2 are chosen to be 12 , providing a smooth roll on and roll off. The 200 H z limit is
chosen such that the weighting filter does not have infinite gain at high frequencies. Finally, with
36

PSD Fy /PSD Fz []

10

10

= 1
= 2
= 3
= 4
= 5
= 10
= 15
= 20
2

10

10

10

10

10

Frequency [Hz]
Figure 3.10: PSD of variation in lateral force divided by PSD of the variation of the vertical force for a smooth road at 15
m/s.

w Fact , the importance of the actuator weighting filter can be expressed. This weighting filter with
w Fact = 1 can be seen in Figure 3.11 together with the other weighting filters.
Suspension travel
The only requirement on suspension travel is that its amplitude is smaller or equal to the amplitude of the passively sprung vehicle. This results in a frequency independent weighting filter for
the suspension travel:
(3.57)
Wz (s) = wz
Here, wz is the gain that determines the importance of the suspension travel, relative to the other
outputs.

37

40

Wzs

Amplitude [dB]

30

20

Wzt
WF act
Wz

10

WU nmod
0

10

20

30
1
10

10

10

10

Frequency [Hz]
Figure 3.11: Normalized weighting filters.

3.5 Summary
The control objectives are minimization of the ISO weighted sprung acceleration for best comfort
and minimization of dynamic tire compression for best comfort. An improvement in comfort can
not be achieved at the wheel hop frequency due to an invariant point in the transfer function at
that frequency. A similar point exists for the suspension travel.
For LQ control it is assumed that the full state is measurable. Using a quadratic criterion the
optimal gains are determined for full state feedback. It is assumed that parameters of the plant
are fully known and that they are fixed.
With robust control the parameters of the plant are not assumed to be fixed. The sprung mass,
tire stiffness and damping are allowed to vary with a certain range. Using weighting filters, synthesis is performed to calculate a controller that is either focussed on comfort or on handling
with constraints on suspension travel and RMS actuator force. As inputs to the controller the
sprung acceleration, unsprung acceleration and suspension travel are used. Given these three
inputs, one output and the weighting filters a state space controller with a minimum order of
twelve is designed.

38

Chapter 4

Analysis of simulation results


With the spring and damper characteristics of the BMW determined in Section 2.1 the performance of the vehicle over a random road can be determined using the quarter car model. This
will be done first in this chapter. Using the weighting matrix, (3.33), defined in Section 3.3 controllers will be developed either focussed on comfort or handling using linear quadratic control.
Since constant parameters are assumed with this control approach and variations of the parameters do occur in practice, controllers will also be developed using robust control. For this the
weighting filters defined in Section 3.4.3 will be used.

4.1 BMW 530i performance on random road


A BMW 530i is used is used as the benchmark vehicle. Given the two road types defined in section 2.3.2, the performance of the BMW can be determined. Tables 4.1 and 4.2 show the sprung
acceleration (z sp ), suspension travel (z p ) and dynamic tire compression (z t p ) of the BMW on a
smooth and rough road respectively, here the subscript p denotes the passive BMW. Figure 4.1
shows the power spectral density of these variables. Visible are the sprung resonance at 1.45 H z
which is caused by the sprung mass resonating on the suspension stiffness. Furthermore, at
13.8 H z the unsprung resonance, called wheel hop frequency, is visible.
Since with the active suspension the suspension properties can be changed, it is interesting
to see what would happen if the spring and damper properties of the BMW are changed. This is
done by scaling the spring and damper characteristics as defined in Chapter 2 while limiting the
suspension travel to that defined by the baseline vehicle. Figure 4.2 shows the results, where it
can be seen that the suspension of the BMW is tuned more to handling than comfort. An improvement of only 2 % can be achieved in dynamic tire compression at the cost of 19 % comfort.

Table 4.1: Performance of the baseline BMW on smooth road.

Parameter
RMS z sp
RMS z swp
Max z p
RMS z t p

Value
0.597
0.492
12.921
1.046

Unit
[m/s 2 ]
[m/s 2 ]
[mm]
[mm]

Description
RMS sprung acceleration
RMS ISO2631 weighted sprung acceleration
Maximum suspension travel
RMS dynamic tire compression

39

Table 4.2: Performance of the baseline BMW on rough road.

PSD(zp )[m2 /Hz]

PSD(
zswp )[m2 /s4 /Hz]

Parameter
RMS z sp
RMS z swp
Max z p
RMS z t p

Value
1.275
1.005
32.935
2.887

Description
RMS sprung acceleration
RMS ISO2631 weighted sprung acceleration
Maximum suspension travel
RMS dynamic tire compression

Weighted sprung acceleration


2

10

10

10

10

10

10

Suspension travel

10

10

10

10

PSD(ztp )[m2 /Hz]

Unit
[m/s 2 ]
[m/s 2 ]
[mm]
[mm]

10

10

10

Dynamic tire compression

10

10

10

10

10

10

10

10

Frequency [Hz]
Figure 4.1: Power spectral density of a BMW 530i. Sprung acceleration, suspension travel and dynamic tire compression
on a smooth road.

If the damping value would be lowered, more comfort would be achieved at the cost of higher
dynamic tire compression. Also visible is that comfort can not be improved indefinitely, this is
caused by the suspension travel reaching its limits. The maximum improvement achievable is
35.3 % at the cost of 40 % increase of the dynamic tire compression. An important note to this
figure is that only one point on the surface can be chosen with the passive system whereas with
an active system one could vary the operating point.

40

zs [m/s 2 ]

zt [mm]

220

220
0.75

200

200

1.65

180

1.6

160

1.55

0.7
180

0.5

80
60

1.45

120

1.4

100

0.45

80

0.4

60

100

1.5

140

1.35

BM
W

0.55

120

BM
W

ds /dsNom [%]

0.6

140

ds /dsNom [%]

0.65

160

1.3
1.25

40

40

0.35

1.2
20

100

20

200

ks /ksNom [%]

100

200

ks /ksNom [%]

Figure 4.2: Sprung acceleration and dynamic tire compression when varying spring stiffness and damping value. Suspension travel of all solutions is smaller or equal to that of the BMW.

41

4.2 Linear quadratic control


In LQ control, design of the controller is done using a quadratic criterion
Z t1
J = lim
ycT (t) Qyc (t) dt,
t1

(4.1)

t0

where the output yc is multiplied by a weighting matrix

q1 0 0
Q = 0 q2 0 .
0 0 q3

(4.2)

With this matrix the outputs, being dynamic tire compression, z t , sprung acceleration, z s , or suspension travel, z, can be emphasized. Given the objectives and constraints defined in Chapter 3,
q1 and q2 are used to either emphasize handling or comfort, whereas q3 is used to emphasize
and thereby limit suspension travel to the maximum value of the BMW. The actuator force limit
of 1000 N is determined by a combination of the three weighting factors. By varying as introduced in (3.3), various controllers can be designed as shown in Figure 4.3. The comfort optimal

300

C
fo
om
rt

250

op
l

150

100

50

50

100

Han
dling o
ptima

B
MW

Ft /Ftp [%]

al
tim

200

150

200

zsw /
zswp [%]
Figure 4.3: Possible controllers achieved with LQ control.

and handling optimal are limited by suspension travel and actuator force respectively. Figure 4.4
shows time domain plots of the handling and comfort optimal controllers. For the comfort optimal controller it is clearly visible that sprung acceleration is minimized. This does result in
high tire deflection as is expected. The actuator signal contains less high frequencies than the
handling optimal controller, this is caused by the sprung acceleration having great influence on
the sprung acceleration. The controller therefore tries to suppress this resonance peak resulting
42

in the low frequencies in the actuator signal. The suspension travel achieved by the comfort controller also clearly contains the same frequency content and is larger than that achieved by the
handling controller.

zsw [m/s2 ]

5
2.5
0
2.5
5
5.5

5.75

6.25

6.5

Handling

0.03

Comfort

z[m]

0.015
0
0.015
0.03
5.5

5.75

6.25

6.5

5.75

6.25

6.5

6.25

6.5

0.03

zt [m]

0.015
0
0.015
0.03
5.5

Fact[N]

5000
2500
0
2500
5000
5.5

5.75

time[s]
Figure 4.4: Time domain plot of sprung acceleration, suspension travel and dynamic tire compression for the maximum
handling and comfort LQ controller on rough road.

As discussed in Section 3.3, LQ control requires the parameters of the model to be linear
and constant. This is obviously not the case in the real car and might give rise to bad performance or instability as Figure 4.5 shows. Here, the controller is designed for a damping value
of 1450 [N s/m] whereas the damping present is only 900 [N s/m], such a situation occurs with
high suspension velocities. As is visible the sprung acceleration increases exponentially and the
controlled system is unstable.
43

25
20
15

zs [m/s 2 ]

10
5
0
5
10
15
20

0.5

1.5

time [s]
Figure 4.5: Sprung acceleration with LQ controller with lower damping than controller was designed for.

4.3 Robust control


As Section 3.4.3 shows, sprung acceleration, dynamic tire compression, actuator force and suspension travel can be influenced by weighting filters. By increasing wz s vertical acceleration can
be emphasized and thereby lowered. However, suspension travel and actuator force have to be
taken into account, this is done by choosing wz and w Fact respectively. Conversely, by increasing
wzt the dynamic tire compression is lowered. The results of this can be seen in Figure 4.6. Limited by suspension travel the comfort can be increased by 60.7 % with a deterioration of a factor
2.2 in dynamic tire compression. The actuator force required by this controller is 501 N . The
handling optimal controller (controller 11) achieves a decrease in the dynamic tire compression
of 21.2 % limited by the maximum RMS actuator force of 1000 N . This controller clearly shows
the effect of the weighting filter for vertical acceleration since the increase in vertical acceleration
is 123 % the ride comfort increase is only increased by 41.8 %.
Furthermore visible in Figure 4.6 are controllers 2-10, these controllers are chosen such that
they are approximately 10 % less comfortable than the previous controller. Controller 5 has the
smallest RMS actuator force, only requiring 134 N RMS with an average power of 10.7 W neglecting losses in the amplifier, see Figure 4.7. This low force requirement is explained by the fact that
the objective of controller 5 is almost equal to the performance of the actuator switched off. The
smallest power requirement is that of controller 6, only requiring 5.5 W , whilst the copper losses
are 5.2 W . The large amount of negative power, i.e. power that flows back to the battery, explains
this low power demand as Figure 4.8 shows.
Controller 7 is close to the BMW specifications. Equally weighting vertical acceleration and
dynamic tire compression in (3.3) ( = 0.5) gives 1 for the BMW and 0.98 for controller 7,
indicating a 2 % increase in overall performance. This small increase is caused by the robust
requirement and the already near to optimal BMW spring and damper. The benefit of active
suspension therefore has to be found in the possibility of switching the controller from a com44

240

Controllers

220

C
om

200

rt
op

160

140

120

H
andl
in g o
p tim
al

al

zt /ztp [%]

tim

A
ctive
stru t
o ff
P
a ssiv
e BM
W

fo

180

100

80

60
20

40

60

80

100

120

140

zsw /
zswp [%]
Figure 4.6: Possible controllers with robust control on rough road, with controller 1 the comfort optimal and controller 11
the handling optimal.

fort objective (controllers 1-6) on straight roads and a handling objective (controllers 7-11) when
cornering.
Robustness
This section has shown that when considering the characteristics of the BMW and actuator as
defined in Chapter 2 good performance is achieved. The two road signals, however, do not guarantee that the full range of possible variations is span. In Section 3.4.2 the structured singular
value, , was introduced to prove robustness. This value has to be smaller than one for the system
to be stable. In Table 4.3 the structured singular value is given for all eleven controllers, there it
can be seen that < 1 for all of them. This worst case structured singular value is achieved with
a damping value of 900 N s/m, tire stiffness of 370e3 N /m and sprung mass of 352.5 kg.

45

1000
900

Fact

Fact [N ] / Ps [W ]

800

Ps

700
600
500
400
300
200
100
0

10

11

Controller
Figure 4.7: Force and power for each of the eleven controllers.

Supply power

150

Copper losses

Ps [W ]

100
50
0
50
100
150
4

4.2

4.4

4.6

4.8

time [s]
Figure 4.8: Supply power and copper losses for controller 6.

4.4 Summary
In this chapter the performance of the BMW as well as the controllers designed with linear
quadratic control and robust control is shown. From the simulations performed with the BMW
spring and damper specifications it was clear that the setup of the car is mostly aimed at handling,
little can be improved by making the dampers stiffer. For both control approaches it is found that
the performance of the comfort optimal controller is limited by suspension travel. For the handling controllers the limiting factor is maximum RMS actuator force. With the linear quadratic
controller, comfort can be improved with 70 %, however, stability can not be guaranteed when
plant parameters are changing. This stability can be guaranteed with robust control as is shown
by the structured singular value that is smaller than one for all controllers. This control approach
allows for the comfort to be improved by 60.7 %.
46

Table 4.3: Maximum structured singular value of all the eleven controllers.

Controller
1
2
3
4
5
6
7
8
9
10
11

[-]
0.3175
0.2657
0.2723
0.2141
0.1924
0.1880
0.2065
0.2233
0.2442
0.2828
0.2982

47

Frequency [Hz]
415
410
412
410
410
612
651
652
654
649
650

Chapter 5

Quarter car test setup


To verify the results obtained with the simulation model, measurements have to be performed.
Due to the many uncertainties in a real car, it is better to start with a quarter car test setup. This
test setup will be introduced in the next section. After this control of the road actuator will be
discussed. Since a problem occurs with one of the sensors, a solution to this problem will be
given. Finally the correlation between measurements and simulations will be shown.

5.1 Description of the test setup


Figure 5.1 shows the full size quarter car test setup. The road disturbances are created by an industrial tubular actuator in parallel with a spring (a) to support the moving weight of the setup.
Control of the actuator is done using standard PID control with notches, this will be discussed
in more detail in Section 5.2. The tire stiffness, kt , is represented by a coil spring (b) and can
be replaced by various springs with different a different stiffness. The unsprung mass, m u , is
represented by a dead weight (c) and is guided by linear bearings only providing a motion in
vertical direction. The sprung mass (e), m s , is connected via the suspension strut (d) to the unsprung mass. The quarter car setup has been designed such that the strut can be easily replaced
by either the active or passive strut. Furthermore, the sprung mass provides the possibility of
adding weight such that parametric uncertainties can be studied. This mass is also guided by
linear bearings only providing freedom in z-direction. Parameters are summarized in Table 5.1.
The benefit of the quarter car test setup is that variables can be measured that can not be

Table 5.1: Test setup parameters.

Parameter
ms
ks
mu
kt
m ra
kra

Value
minimum 340
30.01e3
48.9
352.3e3
11
30.01e3

Unit
[kg]
[N /m]
[kg]
[N /m]
[kg]
[N /m]

48

Description
Sprung mass
Suspension stiffness
Unsprung mass
Tire stiffness
Road actuator mass
Road actuator spring

(e)

(d)
(c)
(b)
(a)

Figure 5.1: Quarter car test setup, with left actual test setup and right schematic representation; (a) road shaker, (b) tire
spring, (c) unsprung mass, (d) suspension strut and (e) sprung mass.

measured easily on a real car. For this purpose various sensors are installed on the test setup.
Starting from the bottom up, the road displacement is measured using an incremental encoder
attached to the tubular actuator. A MicroEpsilon ILD1402-200SC laser sensor is used to measure
the displacement of the unsprung mass. Using this measurement the tire deflection can be
determined
z t = z u zr .
(5.1)
On the unsprung mass, a Kistler 8305B50 accelerometer is fitted. This sensor will also be present
in the real car and will be used as a control input for the robust controller. As Section 2.2 showed,
the suspension travel, z, is of the utmost importance for correct commutation and thus optimal force. This displacement is therefore measured directly by a second MicroEpsilon ILD1402200SC. This sensor, however, has a permanent time delay of 2-2.7 ms [47], a solution for this
delay will be discussed in Section 5.3. This sensor output will also be used as a control input for
the robust controller and can be used to estimate the position of the sprung mass
zs = z + zu .

(5.2)

The last sensor fitted to the quarter car setup is a Kistler 8330A3 accelerometer. This sensor
49

is used to measure the acceleration of the sprung mass and can therefore be directly used to
estimate comfort of the car. This sensor is also a control input for the robust controller. On both
acceleration sensors a second order low pass filter at 200 H z is used to reduce measurement
noise and the influence of test setup resonances, which typically occur above 400 H z.
There is no force sensor installed on the test setup, it is however assumed that the force
determined by the controller is delivered as the current-force relation is included as lookup table
in the test setup. To measure the power that goes to the actuator, three phase to phase voltages as
well as three pase currents are measured as Figure 5.2 shows. To calculate the power supplied to
the actuator, the following equation holds
Ps = Ia Van + Ib Vbn + Ic Vcn ,

(5.3)

since in the shown connection, n, is not available and


Ia + Ib + Ic = 0

(5.4)

Ps = Ia Van + Ib Vbn + (Ia Ib ) Vcn = Ia (Van Vcn ) + Ib (Vbn Vcn ) .

(5.5)

holds, this has to be rewritten to

This finally results in


Ps = Ia Vac + Ib Vbc .

(5.6)

a
Ia
Vab

Vac
n
Ic

b
Vbc

Ib
c

Figure 5.2: Currents and voltages measured in actuator.

5.2 Control of road actuation


As was mentioned before, a PID controller with notches is used to control the road actuator
displacement. Figure 5.3 shows the open loop transfer function from Fra to zr not considering
disturbance forces from Fact . Notable are the 1.45 H z resonance of the sprung mass and the
33.5 H z resonance of the industrial actuator mass, m r . Implementing the controller
Cr oad = 1.3e6
| {z }
Gain

1
20.33
s 2 + 2
s+
1.45
(2 1.45)2
1
s 2 + 20.5
s+1
2
(2 )2

{z

1
s2
(2 33.9)2
1
s2
(2 12)2

}|

N otch1

+
+
{z

20.11
s+1 1 +1
2 33.9
2 4
,
21
1
s
+
1
+1
2 12
2 45

N otch2

50

} | {z }
Lead Filter

(5.7)

Magnitude [dB]

60
80
100
120
140
1
10

10

10

10

Phase [deg]

0
50
100
150
200
1
10

10

10

10

Frequency [Hz]
Figure 5.3: Open loop transfer function from Fra to zr .

Magnitude [dB]

40
20
0
20
1
10

10

10

10

Phase [deg]

150
200
250
300
350
1
10

10

10

10

Frequency [Hz]
Figure 5.4: Open loop controlled transfer function from Fra to zr controlled.

51

results in a open loop bandwidth of 30 H z as Figure 5.4 shows.


Obviously, forces generated by the active suspension strut cause disturbances in zr . To counteract this, the transfer function from Fact to Fra is determined as being
HFactFra =

kt m s s 2
Fact
=
. (5.8)
Fra
ks kt + ds kt s + (kt m s + ks (m s + m u )) s 2 + ds (m s + m u ) s 3 + m s m u s 4

This transfer function is now used as a feed-forward gain for Fra as can be seen in Figure 5.5.
Performance of the system will be discussed in Section 5.4.

Fact
zrref +

HF actF ra
+

error
Croad

+
Road
actuator

zr

Figure 5.5: Block scheme of road control.

5.3 Kalman filter suspension travel


The suspension travel laser sensor, a Micro Epsilon optoNCDT 1402-200, has a permanent delay
of 2 up to 2.7 ms at its highest sample rate of 1500 Hz due to internal calculations being performed. Taking into account the pole pitch of 7.7 mm, a displacement error of 35 % can occur at
1 m/s, resulting in a wrong commutation of the actuator and hence a force that is different from
the force desired. Furthermore, a position error obviously results in an error in the velocity. To
solve this problem, a state observer has to be constructed, since the full quarter car model is nonlinear, a linear model has to be considered. Using the suspension kinematics, the correct position
and velocity can be estimated using the sprung and unsprung acceleration sensor together with
the delayed position measurement. Considering the state vector:
xO =

v z x1 . . . xn

T

(5.9)

where v and z are the velocity and displacement of the suspension, x1 to xn are states resulting
from a Pad approximation [22] of the time delay. Using a second order Pad approximation,
which has a matching phase up to 150 H z for a 2.7 ms delay, the state space systems reads:



2222 1608
64
xd =
xd +
z
1024
0
0
|
{z
}
| {z }


Ad

zd =

(5.10)

Bd

 

69.44 0 xd + 1 z.
{z
}
| {z }
|


Cd

Dd

52

(5.11)

Together with the state vector, (5.9), this results in:

v
0 0
0 0
1 1 


z 1 0

0 0
x O + 0 0 as
=

x1 0 Bd

0 0 au
Ad
x2
0 0
0 0


z d = 0 Dd C d x O

(5.12)

(5.13)

With as and au the acceleration of the sprung and unsprung mass respectively. Observability of
the systems is determined by checking if
h
CT

AT C T

AT

2

CT

...

AT

n1

CT

iT

= n,

(5.14)

with n the size of the state, the system is observable. Here, with n = 4 this is the case. Keeping
in mind the standard plant form for designing a Kalman filter equals
x = Ax + Bu + Gw

(5.15)

y = C x + Du + H w + v

(5.16)

with w the white process noise and v the measurement noise. Given that


E ww T = Q and E vv T = R

(5.17)

the optimal Kalman filter solution can be calculated with equations



x = A x + Bu + L y C x Du

(5.18)

The filter gain L is found solving the Riccati equation. Solving this results in a Kalman filter
that is capable of estimating the state and thereby the real suspension travel, as can be seen in
Figure 5.6. This result has been achieved by experimentally determining the Q and R resulting
in


50 0
Q = 0.005
,
(5.19)
0 1
R = 1e7.

(5.20)

The maximum estimation error is 0.45 mm, whereas the maximum error of the laser sensor
would be 0.81 mm. When looking at the RMS improvement in error, a similar improvement can
be seen. Without Kalman filter the RMS error is 0.2 mm, with the Kalman filter the RMS error is
0.107 mm.

5.4 Experimental validation of setup


To validate the test setup, the road displacement is used as an input to the system. As power
spectral density of the measured and simulated quarter car test setup show in Figure 5.7, the
road input signal is followed up to 30 H z, as could be expected from the bandwidth of the road
actuator controller. As can be seen from the ISO2631-1 weighting criterion, humans are most
sensitive between 4 H z and 10 H z which is below the 30 H z road excitation limits, therefore, this
tracking is sufficient. The sprung acceleration, unsprung acceleration and suspension travel also
53

x 10
4

z [m]

2
0
True value
Kalman estimate

Sensor value
4
2.9

2.95

3.05

3.1

3.15

3.2

3.25

3.3

3.35

3.4

x 10

Kalman error
Sensor error

Error z [m]

4
2
0
2
4
2.9

2.95

3.05

3.1

3.15

3.2

3.25

3.3

3.35

3.4

time [s]
Figure 5.6: Measurement versus Kalman estimate of real position on smooth road.

match closely up to 30 H z. This is also verified by Table 5.2 where the RMS acceleration levels
and suspension travel are shown. The small differences are caused by friction in the bearings
and setup misalignment. Furthermore, transmission of vibrations through the construction of
the test setup causes higher measured accelerations. Additionally, a tire spring with a constant
stiffness is used in the test setup. Finally, cogging in the electromagnetic tubular actuator results
in the suspension compressing differently than when the motion would be fluent. These effects
are, however, considered to be small enough to ignore in the model. Figure 5.7 also shows resonances at 64.5 H z and 141 H z, these resonances of the support spring parallel to the industrial
actuator, kra , and the tire spring respectively. From this it can be concluded that the simulations
and measurements closely match.

Table 5.2: Comparison of RMS values of simulation and measurement on quarter car test setup using a smooth road.

z s
zt
z

Measured
0.623 m/s 2
0.974 mm
9.749 mm

Simulated
0.597 m/s 2
1.046 mm
12.92 mm

54

Description
Sprung mass acceleration
Dynamic tire compression
Suspension travel

PSD zr [m2 /Hz]


2
4
PSD zu [m2 /s 4 /Hz] PSD zs [m /s /Hz]

PSD z [m2 /Hz]

Road displacement

10

10

Measured

10

Simulated
15

10

10

10

10

10

Sprung acceleration

10

10

10

10

10

10

10

10

Unsprung acceleration

10

10

10

10

10

10

10

Suspension travel

10

10

10

15

10

10

10

10

10

Frequency [Hz]

Figure 5.7: Power spectral density of road signal, sprung acceleration, unsprung acceleration and suspension travel,
measurement and simulation compared.

5.5 Summary
In this chapter the full size quarter car test setup is introduced. It consists of an industrial actuator in parallel with a spring. This combination is used to put road disturbances on the quarter
55

car. The tire stiffness is represented by a second spring attached to a block of steel that represents
the unsprung mass. The unsprung mass is connected to the sprung mass via the suspension
strut. Various signals are measured on the test setup, being the sprung and unsprung acceleration, suspension travel, tire and road deflection. For control of the industrial actuator a PID
controller with notches is used, furthermore, a feed forward term is introduced that accounts for
disturbances caused by the active suspension.
To account for sensor delay, a Kalman filter is designed with which, using the two acceleration
sensors, the position is predicted. The Kalman filter is designed using a Pad approximation of
the delay. The RMS position error is reduced approximately 50 % compared to the delayed signal
using the Kalman filter.
Experimental validation of the setup showed that all signals match closely up to 30 Hz. Beyond that frequency, the industrial actuator is not capable of following the road reference signal.
Two resonances can be observed at 64.5 H z and 141 H z, these are caused by resonance of the
support spring and tire spring respectively.

56

Chapter 6

Measurement results achieved on


quarter car setup
Using the quarter car test setup as defined in the previous chapter together with the controllers as
shown in Chapter 4, measurements can be performed to determine the quality of the controllers.
First, results will be shown using two linear quadratic controls on random road, one focussed on
comfort, the other focussed on handling. After this the performance achieved with the eleven
robust controllers on random road will presented. Finally, the most comfortable controller will
be used to perform measurements with a speed bump as disturbance signal.

6.1 Linear quadratic control


Figure 6.1 shows the results obtained with LQ control for the comfort and handling optimal controller as were introduced in Section 4.2. As is visible for the comfort on the smooth road, the
deviation from the simulated value is 55 % which is most likely caused by friction, cogging, non
linearities in the test setup and errors in the measured state. Nonetheless a 35 % improvement in
comfort is achieved. The deviation from the simulation can also be observed in dynamic tire compression for the comfort controller. The tradeoff between comfort and handling does, however,
still hold because the dynamic tire compression is 34.1 % lower than predicted in simulation.
The measured value of the sprung acceleration for the handling controller on the smooth
road does match the simulated value better. This also holds for the dynamic tire compression.
As is visible, an improvement of 48.5 % in dynamic tire compression is achieved at the cost of
23 % in comfort. This is also true for the suspension travel, however, not for the actuator force
which is 33 % higher. This is confirmed by the PSD of the actuator force as Figure 6.2 shows.
Between 3 and 5 H z this clearly higher as well as at 30 H z. This 30 H z deviation is caused by a
resonance of the road actuator. Since the full state, including zr is measured and used for control,
this causes the actuator force to be larger at this frequency.
The same trend that was visible for the comfort controller on smooth road is visible for the
comfort controller on 50 % rough road. The road level has been changed because the road displacement actuator was not capable of delivering more RMS force when the active suspension
was in operation. Furthermore it is visible that measurements of the handling controller on
rough road have not been performed, this is due to the controller becoming unstable.

57

RMS zsw [m/s 2 ]

Smooth road

50% Rough road

(a)
1

0.5

RMS zt [mm]

(b)
4
3
2
1
0

Max z [mm]

(c)
20
15
10
5
0

RMS Fact [N ]

(d)
1000

500

Mean Ps [W ]

(e)
150
Measured
Simulated

100
50
0

Comfort

Handling

BMW

Comfort

Handling

BMW

Figure 6.1: Measured vs simulated performance of LQ comfort and handling controller with, (a) RMS weighted sprung
acceleration, (b) RMS dynamic tire compression, (c) maximum suspension travel, (d) RMS actuator force and (e) supply
power.

58

PSD
zsw [m2 /s4 /Hz]
PSD z[m2 /Hz]

Weighted sprung acceleration


3

10

10

10
Suspension travel

13

10

10
Dynamic tire compression

10

10

12

10

10

PSD zr [m2 /Hz]

10

10

10

PSD zt [m2 /Hz]

10

10
Road displacement

10

10

12

10

Simulated
Measured
Measured BMW
0

10

PSD Fact[N]

10
Actuator Force

10

10

10

10

10

10

10

Frequency [Hz]

Figure 6.2: Power spectral density of sprung acceleration, suspension travel, dynamic tire compression, road displacement
and actuator force for the LQ handling controller on smooth road.

59

6.2 Robust control


The controllers that have been introduced in Section 4.3 have been tested on the quarter car test
setup. Figure 6.3 and Figure 6.9 show the results of this. Due to RMS force limitations of the
industrial actuator, the rough road could not be turned on fully when the suspension actuator
was in operation. Therefore the rough road presented here will be 60 % of the road presented in
Section 2.3.2, the gain will thus be 0.075 instead of 0.125.
The best comfort on the smooth road is achieved by controller 2 as Figure 6.3 (a) shows.
Compared to the measured BMW suspension a performance gain of 48 % is achieved at the
cost of 99.3 % dynamic tire compression. The fact that measurement deviates 63.8 % from the
simulation is explained partially by friction in the linear bearings, but mostly by stick slip behavior
of the actuator. This is corroborated by the difference in sprung acceleration of the simulated and
measured actuator which differs 46.9 %. Furthermore, transmission of vibrations from the road
actuator to the sprung mass via the measurement frame also influence the measured acceleration.
It has to be noted that the actuator force predicted by the simulation is 60 % lower than in the
measurement. This higher actuator force also clearly has its influence on the actuator power as
Figure 6.3 (e) shows. This higher actuator force is mostly caused by deviations from the simulated
actuator force between 3.5 and 10 H z and 15 to 45 H z as Figure 6.4 shows.
Figure 6.4 also clearly shows the effect of the frequency dependent weighting filters. In the
weighted sprung acceleration, the largest reduction is achieved between 4 and 10 H z. In the
actuator force, a change of slope is clearly visible beyond 30 H z which is what was desired by the
weighting filter.
The fact that controller 1 is less comfortable than controller 2 in Figure 6.3 (a) is explained
by stick slip behavior of the actuator. When measuring the stick slip behavior of the suspension
actuator it is found that minimally 70 N is necessary to break this friction in the positive force
direction and 90 N in the negative force direction as Figure 6.6 shows. Modeling this behavior as
proposed by Wild [48] and incorporating this in the simulations results in the results presented
in Figure 6.7. It is now clearly visible that controller 1 is also less comfortable in simulation than
in controller 2. This is also true for controller 5 which was also found to be less comfortable than
controller 6 during measurements. Concluding from this is that incorporating static friction in
the design of the controllers is of importance to obtain correct results.
In terms of handling, a performance gain of 17.7 % is achieved, worsening comfort by 10.7 %
for controller 11. The lower value measured compared to simulation is primarily caused by friction
in the bearings of the unsprung mass. This results in a smaller displacement of the unsprung
mass but also in more transmission of vibrations to the sprung mass. The measured RMS actuator force and power are similar to those predicted in the simulations. Figure 6.5 shows dynamic
tire compression of controller 11 versus that of the BMW. As is visible, the tire compression is
usually smaller when controlled explaining the improvement in dynamic tire load.
The difference in actuator force as is most clearly visible for controllers 6, 7 and 8 is primarily
caused by the difference in sprung acceleration. Due to friction in the bearings of the sprung
mass and actuator this value is higher, the controller wants to correct this resulting in higher
actuator forces. This is supported by the power spectral density of the sprung acceleration and
actuator force of controller 7 as is shown in Figure 6.8.

60

RMS zsw [m/s 2 ]

(a)
1

0.5

RMS zt [mm]

(b)
2
1.5
1
0.5
0

Max z [mm]

(c)
15
10
5
0

RMS Fact [N ]

(d)
400
300
200
100
0

Mean Ps [W ]

(e)
60
Measured
40

Simulated

20
0

10

11

BMW Actuator off

Controller
Figure 6.3: Measured vs simulated performance of robust controllers on smooth road with, (a) RMS weighted sprung
acceleration, (b) RMS dynamic tire compression, (c) maximum suspension travel, (d) RMS actuator force and (e) supply
power.

61

PSD z[m2 /Hz] PSD zsw [m2 /s4 /Hz]

Measured
Weighted sprung acceleration
10

10

10

PSD zt [m2 /Hz]

10
Suspension travel

10

10

13

10

10
Dynamic tire compression

10

10

12

10

10

PSD zr [m2 /Hz]

Passive Measured

10

10
Road displacement

10

10

10

10

10
5

PSD Fact[N]

Simulated

10
Actuator Force

10

10

10

10

10

10

Frequency [Hz]

Figure 6.4: Power spectral density of sprung acceleration, suspension travel, dynamic tire compression, road displacement
and actuator force for controller 2 on smooth road.

62

x 10
4

Active
3

BMW

0
1
2
3
4

5.1

5.2

5.3

5.4

5.5

5.6

5.7

5.8

5.9

time [s]
Figure 6.5: Dynamic tire compression achieved by controller 11 versus BMW.

150

100

50

Force [N ]

zt [m]

Static friction

Measured force
50

100

150

200
0.05

0.05

z [m]
Figure 6.6: Stick slip behavior of actuator measured at 0.025 m/s.

63

0.7
Measured

RMS zsw [m/s 2 ]

0.6

Simulated

0.5
0.4
0.3
0.2
0.1
0

10

11

BMW Actuator

Controller
Figure 6.7: Weighted sprung acceleration on smooth road with robust controller, measured vs simulated. Stick slip model
used with slip boundary of 90 N .

This deviation caused by the disturbances has less influence for the rough road measurements, because their relative level compared to the amplitude of the sprung acceleration is lower
as Figure 6.9 (d) shows. The correlation between simulations and measurements is therefore
more consistent. This is also valid for the sprung acceleration and dynamic tire compression.
Again, controller 2 is the most comfortable controller with a comfort improvement of 40 % resulting in a deterioration in dynamic tire compression of 83.6 %.
The handling optimum is achieved by controller 11 as can be seen in Figure 6.9 (b). A 25.5 %
improvement is reached, worsening comfort by only 6 %. This smalle decrease clearly shows the
effect of the weighting filters again since sprung acceleration is worsened by 75 % whereas the
weighted sprung acceleration was only decreased by 6 %. Furthermore, maximum suspension
travel is slightly smaller than that of the BMW.
A noticeable difference between simulated and measured actuator power occurs for controllers 6, 7 and 8, this is partially caused by the higher actuator force and partially caused by
the higher suspension speeds as Figure 6.10 shows. Higher velocities peaks can be observed
which result in more power.
Driving over the speed bump as introduced in Section 2.3.2 leads to the results as shown in
Figure 6.11 when using controller 1. As is visible maximum sprung acceleration, z s , is reduced by
53.3 % compared to the BMW. This improvement does cost some suspension travel as both the
simulation and measurement show. The maximum absolute actuator force required is 944 N
with a maximum power requirement of 883.6 W when driving down the bump again. The noise
that can be observed on the measured sprung acceleration is caused by a setup resonance at
455 H z.
64

PSD zsw [m2 /s4 /Hz]


PSD z[m2 /Hz]

Weighted sprung acceleration


3

10

10

10
Suspension travel

10

10

10
Dynamic tire compression

10

10

12

10

10

PSD zr [m2 /Hz]

10

10

10

PSD zt [m2 /Hz]

10

10
Road displacement

10

10

10

10

15

10

Measured
Simulated
Passive Measured
0

10

PSD Fact[N]

10
Actuator Force

10

10

10

10

10

10

Frequency [Hz]

Figure 6.8: Power spectral density of sprung acceleration, suspension travel, dynamic tire compression, road displacement
and actuator force for controller 7 on smooth road.

Robustness
In simulations (Section 4.3), the structured singular value was used to show that all controllers
are robust given the variations present in the plant. Determining this value in the test setup
is impossible, measurements will therefore be performed with the worst case parameters. This
65

RMS zsw [m/s 2 ]

(a)
1

0.5

RMS zt [mm]

(b)
4
3
2
1
0

Max z [mm]

(c)
30
20
10
0

RMS Fact [N ]
Mean Ps [W ]

(d)
600

150

400
200
0

(e)
Measured

100

Simulated

50
0

10

11

BMW Actuator off

Controller
Figure 6.9: Measured vs simulated performance of robust controllers on rough road with, (a) RMS weighted sprung
acceleration, (b) RMS dynamic tire compression, (c) maximum suspension travel, (d) RMS actuator force and (e) supply
power.

66

0.4
Measured
0.3

Simulated

v [m/s]

0.2
0.1
0
0.1
0.2
0.3
0.4

2.1

2.2

2.3

2.4

2.5

2.6

2.7

2.8

2.9

time [s]
Figure 6.10: Measured versus simulated suspension velocity with controller 8.

means that the tire stiffness is chosen as stiff as possible (352.3e3 N /m) and a sprung mass that
is as light as the minimum weight of the car (352.5 kg). The damping value can not be altered,
but will vary constantly when in operation. Given the unmodeled dynamics of the test setup and
the worst case tire stiffness and sprung mass, it is assumed that if the controller is stable for this
situation, the controller is stable for each situation allowed. As Figure 6.12 shows, controller 1 still
performs better despite the worst case setup parameters. This is also true for all other controllers,
robustness is thereby proven. Performance is also better compared to the BMW for this controller.

6.3 Summary
In this chapter results obtained on the quarter car test setup using linear quadratic control and
robust control are presented. With linear quadratic control comfort can be improved by 35 %,
whereas handling can be improved by 48.5 %. Deviations from the simulations are caused by
non-linearities in the test setup as well as errors in the measured state.
When using robust control comfort can be improved by 48 %. With the handling controller
an improvement of 17.7 % can be achieved. Deviations from the simulations are largely caused
by static friction of the active suspension. Incorporating static friction in the simulations showed
similar results as the measurements. Frequency dependent weighting filters are clearly effective, as weighted sprung acceleration is only increased by 6 % whereas the non-weighted sprung
acceleration is increased by 75 % for the handling optimal controller.
Implementing the three centimeter high speed bump in the quarter car test setup showed that
sprung acceleration can be decreased by 53.3 % compared to the BMW. Maximum force required
for this is 944 N .

67

zr [m]

0.04
BMW
Simulated
Measured

0.02
0
0.02
0.5

1.5

2.5

1.5

2.5

1.5

2.5

1.5

2.5

z [m]

0.04
0.02
0
0.02

zs [m/s 2 ]

0.04
0.5

2
0
2
0.5

Fact [N ]

1000
0
1000
0.5

time [s]
Figure 6.11: Measured versus simulated road displacement, suspension travel, sprung acceleration and actuator force
when driving over a speed bump for controller 1 (best comfort).

BMW
Controlled

zs [m/s 2 ]

2
1
0
1
2
10

10.2

10.4

10.6

10.8

11

time [s]
Figure 6.12: Controlled sprung acceleration compared with BMW performance for worst case parameters with controller 1.

68

Chapter 7

Conclusions and recommendations


In this research, control of a tubular direct drive electromagnetic active suspension for a BMW 530i
is discussed. The goal is to influence the comfort and handling of the vehicle by means of proper
control of the active suspension, the main research question therefore is:
What performance gains in comfort and handling can be achieved with a high bandwidth electromagnetic active suspension given the constraints of maximum actuator force and suspension travel?
This chapter will draw conclusions upon the research question and objectives. Furthermore,
recommendations will be formulated that can be incorporated in future research.

7.1 Conclusions
The main conclusion of this report is that robust control offers the best possibility of controlling the active suspension. Simulations show that an improvement of 60.7 % in comfort can
be achieved deteriorating dynamic tire compression by 125.7 %. Furthermore, when choosing
another controller that is focussed on handling, this can be improved by 21.2 % deteriorating
comfort by 41.8 %. Further improvements are limited by suspension travel for the comfort controller and maximum RMS actuator force for the handling controller. In Table 7.1 a summary of
the main results of this thesis can be found.
Table 7.1: Maximum improvement of simulated and measured performance of LQ and robust controllers (positive is better).

LQ max comfort
LQ max handling
Robust max comfort
Robust max handling

z sw Simulated
73.7 %
-28.9 %
60.7 %
-41.8 %

z sw Measured
35 %
-23 %
48 %
-10 %

z t Simulated
-194.1 %
59.6 %
-125.7 %
21.2 %

z t Measured
-104.3 %
48.5 %
-99.4 %
17.7 %

For controller design a simple model of the BMW has to be used, a two DOF quarter car model
is therefore chosen. As input to this model, first-order filtered white noise was used representing,
depending on the gain, a smooth or rough road. The vertical acceleration of the sprung mass, z s ,
is a good indication of comfort. Humans are most sensitive to vibrations between 4 and 10 H z,
this is properly described by the ISO2631-1 standard. Secondly, the dynamic tire compression, z t ,
is used to asses handling of the vehicle. As constraints for the controller design suspension
69

travel, z, is limited to the suspension travel achieved by the BMW. The second constraint is determined by thermal limits of the actuator, the maximum RMS actuator force therefore is 1000 N .
To test the controllers developed, a full size quarter car test setup is created. In this setup, the
sprung and unsprung mass are represented by blocks of steel and are guided by linear bearings.
These two masses are coupled by a suspension strut, which can be replaced easily. The tire stiffness is represented by a coil spring with a nominal stiffness equal to that of the tire. Finally, road
disturbances are provided by a tubular industrial actuator in parallel with a spring to support the
setup weight. In total five sensors are fitted to the test setup from which three are used for control
of the active suspension. The sensors installed are an incremental encoder, measuring the road
displacement, a laser sensor measuring the unsprung displacement, a laser sensor measuring
the suspension travel and two acceleration sensors, measuring the vertical acceleration of the two
masses. The latter three sensors are used as controller inputs for the robust controller.
Two control approaches are considered, linear quadratic control and robust control. In the
LQ control approach, an optimal controller is found, depending on three weighting gains. This
way, either comfort or handling can be emphasized with constraints on suspension travel and
maximum RMS actuator force. A linear model has to be used for this control topology, thereby
neglecting variations in the plant. Robust control does account for these variations. Emphasis on
either comfort or handling for this control topology is done by frequency dependent weighting
filters such as an approximation of the ISO2631-1 criterion for comfort.
By varying the weighting filters, either comfort or handling is emphasized. In total eleven
controllers are designed, with each controller 10 % less comfortable than the previous. The performance of the comfort optimal controlled is limited by suspension travel, whereas the performance of the handling optimal controller is limited by maximum RMS actuator force. For
the comfort optimal controller, the actuator has to deliver 501 N RMS resulting in a power consumption of approximately 300 W . Lower power consumption is achieved by the controllers with
performance close to that of the actuator in passive operation. The power distribution is such that
the average mechanical power requirement is zero, with the only power requirement the copper
losses.
Measurements on the test setup show that, although there is a difference between measured
and simulated values, a clear trend can be observed for each robust controller. The objective
clearly goes from comfort (controller 1) to handling (controller 11) with increasing controller number. A 48 % improvement in comfort can be achieved compared to the BMW. For the handling
case this is 17.7 %, worsening comfort by only 10 %. Deviations from the simulated value are
explained by friction in the test setup, as well as in the tubular actuator. Furthermore, transmission of vibrations through the test setup also makes the acceleration values larger. Measurements
on the rough road show a better correlation between simulation and measurements, proving this
theory due to the lower relative level of the friction.
The LQ controller is designed with constant parameters, this results in stability issues when
measuring the handling controller on the rough road due to the parameters of the setup varying.
For robust control, stability is proven by the structured singular value, which is smaller than one
for all eleven controllers. Measurements also shown that all robust controllers are stable on the
test setup.
Finally, when plotting the controller response in frequency domain, a clear improvement in
comfort can be observed in the 4-10 H z region proving that frequency weighting has its influence.
This also holds for actuator force, which clearly decreases beyond its cut-off frequency.
70

7.2 Recommendations
This thesis tries to predict the performance of an active suspension system by means of a quarter
car model. This is clearly a simplification of a full car model and does not cover all aspects of a
full car such as pitch, roll and yaw behavior. To achieve optimal results in these fields as well, a
more detailed model should be considered that also incorporates these degrees of freedom.
Improvements in the test setup such as reduction of friction in the bearings would improve
the quality of the measurements and thereby also the quality of control. A suggestion for reducing
the friction is changing the size of the sprung mass, thereby reducing tilting of this mass which
results in extra friction. If the block would be chosen small enough (a 0.37 m square cube would
be sufficient), vertical guidance might be omitted. A second improvement in the test setup would
be to increase the capabilities of the road actuator. The maximum RMS force is too small for it to
be capable of running the rough road when the actuator is in operation.
A suspension travel sensor without time delay would improve the quality of control and commutation of the actuator. Although the Kalman filter estimates the suspension travel quite accurately, the low pass property of this filter prevents higher frequency movements.
Using the damping as a known variable would improve the performance of the robust controller since less uncertainties would be included. Given the goniometric nature of the fit of the
damping value, it would be possible using input-output linearization. Another remark about the
fail-safe damping value is that if a smaller value would be chosen, regeneration of energy would
be possible as literature showed. Decreasing damping would have a slight effect on the possible
improvements in dynamic tire compression, as it would require more actuator force to achieve
the same level of performance.
A bettering bearing system would most likely reduce the amount of static friction and thereby
increase the performance. Suggestions for this are a different type of linear bearing or using two
bearings as a normal MacPherson suspension strut does.
Finally, as eleven robust controllers have been proposed, it is possible to switch between them
and change the performance of the vehicle from comfort to handling. It has however not been
proven that this is possible. Literature offers various proofs for minimal switching time, implementing this in a vehicle state controller, that decides whether comfort or handling should be
emphasized would increase usability when implementing the system in a real car.

71

Bibliography
[1] Business technology; future shocks: a new auto suspension system. New York Times,
January 21, 1987.
[2] Mercedes ABC, http://500sec.com/abc-active-body-control. Online, 2010.
[3] Citron hydractive 3, http://www.citroenet.com. Online, 2010.
[4] Delphi magneride semi-active suspension, http://delphi.com. Online, 2010.
[5] Magnetic Marelli variable damping, http://www.magnetimarelli.com. Online, 2010.
[6] P. J. Venhovens, Optimal control of vehicle suspensions. PhD thesis, Technical University Delft,
1994.
[7] L. Power, GenShock by Levant Power. Presentation, 2010.
[8] M. Mnster, U. Mair, H. Gilsdorf, A. Thoma, Mller, M. Hippe, and J. Hoffmann, Electromechanical active body control, ATZ worldwide, pp. 4449, Sep 2009.
[9] W. Jones, Easy ride: Bose corp. uses speaker technology to give cars adaptive suspension,
Spectrum, IEEE, vol. 42, pp. 12 14, May 2005.
[10] Bose Corp. active suspension system, http://www.bose.com. Online, 2010.
[11] B. Gysen, J. Paulides, J. Janssen, and E. Lomonova, Active electromagnetic suspension system for improved vehicle dynamics, Vehicular Technology, IEEE Transactions on, vol. 59,
pp. 1156 1163, March 2010.
[12] B. Gysen, T. Van der Sande, J. Paulides, and E. Lomonova, Efficiency of a regenerative
direct-drive electrmomagnetic active suspension, VPPC conference, 2010.
[13] K. Eckert, Suspension system for a car. Patent, 1973.
[14] T. Dahlberg, An optimized speed-controlled suspension of a 2-DOF vehicle travelling on a
randomly profiled road, Journal of sound and vibration, vol. 62, no. 4, pp. 541546, 1978.
[15] P. Michelberg, L. Palkovics, and J. Bokor, Robust design of active suspension system, Int.
J of Vehicle Design, vol. 14, no. 2/3, pp. 145165, 1993.
[16] A. de Jager, Comparison of the two methods for the design of active suspension systems,
Optimal control applications and methods, vol. 12, pp. 172188, 1991.
72

[17] M. Yamashita, K. Fujimori, K. Hayakawa, and H. Kimura, Application of h control to active


suspension system, Automatica, vol. 30, no. 11, pp. 17171729, 1994.
[18] A. Sharkawy, Fuzzy and adaptive fuzzy control for the automobiles active suspension system, Vehicle system dynamics, vol. 43, no. 11, pp. 795806, 2005.
[19] D. Hrovat, Survey of advanced suspension developments and related optimal control applications, Automatica, vol. 33, no. 10, pp. 17811817, 1997.
[20] J. Doyle, Guaranteed margins for LQG regulators, IEEE transactions on automatic control,
vol. 23, pp. 756757, August 1978.
[21] C. Lauwerys, J. Swevers, and P. Sas, Robust linear control of an active suspension on a
quarter car test-rig, Control engineering practive, vol. 13, pp. 577586, 2005.
[22] C. Lauwerys, J. Swevers, and P. Sas, Design and experimental validation of a linear robust
controller for an active suspension of a quarter car, Proceedings of the 2004 American control
conference, vol. 2, pp. 14811486, 2004.
[23] S. Lee and W. Kim, Active suspension control with direct-drive tubular linear brushless
permanent-magnet motor, IEEE transactions on control system technology, vol. 18, pp. 859
870, July 2010.
[24] BMW, BMW 5-series catalogue, 2010.
[25] J. Janssen, Design of active suspension using electromagnetic devices, Masters thesis,
Eindhoven University of Technology, 2006.
[26] D. Bastow, G. Howard, and J. Whitehead, Car Suspension and Handling. Professional Engineering Publishing, 2004.
[27] R. Blom, J. Vissers, L. Merkx, and M. Pinxteren, Flat plank tyre tester, users manual, tech.
rep., Eindhoven University of Technology, 2009.
[28] B. Gysen, J. Jansen, J. Paulides, and E. Lomonova, Design aspects of an active electromagnetic suspension system for automotive applications, Industry Applications, IEEE transactions on, vol. 45, pp. 15891597, September-October 2009.
[29] B. Gysen, J. Paulides, L. Encica, and E. Lomonova, Slotted tubular permanent magnet actuator for active suspension systems, The 7th International symposium on linear drives for
industry applications, pp. 292295, 2009.
[30] J. Wang, W. Wang, A. K, and H. D, Compartive studies of linear permanent magnet motor
topologies for active vehicle suspension, IEEE vehicle power and propulsion conference, pp. 1
6, 2008.
[31] M. van de Wal and B. de Jager, Acuator and sensor selection for an active vehicle suspension
aimed at robust performance, Int. Journal Coutrol, vol. 79, no. 5, pp. 70.720, 1998.
[32] A. Kruczek and A. Stribrsky, eds., A Full-car Model for Active Suspension - Some Practical
Aspects, Proceedings of IEEE International Conference on Mechatronics, 2004.
73

[33] I. Besselink and A. Schmeitz, Vertical dynamics 4L150, tech. rep., Eindhoven University
of Technology.
[34] ISO, ISO2631-1:1997:Mechanical vibration and shock - Evaluation of human exposure to
whole-body vibration, tech. rep., International Organization for Standardization, Geneva Switzerland, 1997.
[35] C. Dodds and J. Robson, The description of road surface roughness, Journal of sound and
vibration, vol. 31, no. 2, pp. 175183, 1973.
[36] ISO, ISO8608:1995:Mechanical vibration-Road surface profiles-Reporting of measured
data, tech. rep., International Organization for Standardization, Geneva - Switzerland, 1995.
[37] L. Zuo and S. Nayfeh, Low order continuous-time filters for approximation of the iso2631-1
human vibration sensitivity weightings, Journal of sound and vibration, vol. 265, pp. 459
465, 11 2002.
[38] J. Hedrick and T. Butsen, Invariant properties of automotive suspensions, Proceedings of
institution of mechanical engineers, vol. 204, pp. 2127, 1990.
[39] D. Karnopp, How significant are transfer function relations and invariant points for a quarter car suspension model?, Vehicle system dynamics, vol. 47, no. 4, pp. 457464, 2009.
[40] H. Kwakernaak and R. Sivan, Linear Optimal Control Systems. Wiley, 1972.
[41] S. Skogestad and I. Postlethwaite, Multivariable feedback control. Wiley, 2007.
[42] A. Damen and S. Weiland, Robust control. Lecture Notes with robust control course TUe,
2001.
[43] H. Pacejka, Tyre and vehicle dynamics. Elsevier, 2nd ed., 2006.
[44] M. ElMadany and Z. Abduljabbar, Linear quadratic gaussian control of a quarter-car suspension, Vehicle system dynamics, vol. 32, no. 6, pp. 479497, 1999.
[45] H. Pacejka and I. Besselink, Magic formula tire model with transient properties, Vehicle
system dynamics, vol. 27, pp. 234249, 1997.
[46] D. Anderson and B. Mills, Dynamic analysis of a car chassis fram using the finite element
method, International Journal mechanical science, vol. 14, pp. 799808, 7 1972.
[47] Micro-Epsilon MessTechnik, Instruction Manual optoNCDT 1402, tech. rep., 2008.
[48] H. Wild and S. Dodds, Real-time identification of the friction coefficient of a rolling guided
high dynamic linear motor, International conference on control, 1998.

74

Appendix A

Tire Model
Stiffness and lateral force measurements have been performed on a Dunlop SP Sport
225/50/R17 94W tire. This tire was fitted to a BMW rim with a width of 7.5 inch and a diameter of 17 inch. Its ET value was 20 mm. This run flat tire is normally mounted under a BMW
530i. All measurements are performed on a flatplank tire test present at Eindhoven University of Technology. First of all the vertical stiffness of the tire will be estimated, secondly relaxation measurements will be done, which give an idea of the lateral force build up of the tire. The
Magic Formula (MF) tire model will then be discussed, and the relaxation measurements will be
used to determine the MF parameters.

A.1 Vertical stiffness


The vertical stiffness of a tire is given as:
d Fz
(A.1)
du
where F is the vertical force and u is the tire deformation. Since a standing tire differs from
a rolling tire, measurements are performed while the tire is rolling with a velocity of 5 cm/s.
Measurements have been performed for two tire pressures: 2.4 and 2.8 bar . The measurement
procedure consists of compressing the tire up to approximately 2.4 cm, which corresponds to
roughly 8500 N for the higher tire pressure. Figure A.1 shows the results of this procedure. A
loop can be clearly identified, this is caused by the damping that is present in the tire. It is also
clearly visible that the higher pressure results in a higher maximum force. Furthermore it has to
be noted that the relation between force and deformation is quadratic. Two fits have been made
for this
Fz24 = 2.83e6u 2 + 2.73e5u,
(A.2)
kt z =

Fz28 = 2.92e6u 2 + 2.79e5u.

(A.3)

From these fits the stiffness can be easily derived


kt z24 = 5.66e6u + 2.73e5,

(A.4)

kt z28 = 5.84e6u + 2.79e5.

(A.5)

Figure A.2 shows the stiffness as a function of vertical load. A difference in stiffness of 9000 N/m
can be observed for the different tire pressures. For the nominal case, a load of approximately
4000 N , the stiffness is 346000 N /m for 2.4 bar and 354000 N /m for 2.8 bar .
75

9000
Measurement 1 2.4 bar
Measurement 2 2.4 bar
Measurement 1 2.8 bar
Measurement 2 2.8 bar

Vertical tire force [N ]

8000
7000
6000
5000
4000
3000
2000
1000
0

0.005

0.01

0.015

0.02

0.025

Deformation [m]
Figure A.1: Vertical force as a function of deformation for different tire pressures.

x 10
4.2

2.4 bar
2.8 bar

Vertical stiffness [N/m]

4
3.8
3.6
3.4
3.2
3
2.8
2.6

1000

2000

3000

4000

5000

6000

7000

8000

Vertical tire force [N ]


Figure A.2: Vertical stiffness as a function of vertical force for different tire pressures.

76

500

1 Meas

1 Fit
0

3 Meas
3 Fit
5 Meas

500

5 Fit
7 Meas
7 Fit

Fy [N ]

1000

9 Meas
9 Fit
11 Meas

1500

11 Fit
2000

2500

3000

3500

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Displacement [m]
Figure A.3: Lateral force development for various side slip angles including 1st order fit.

A.2 Relaxation measurements


A tire develops lateral force as a function of side slip angle and vertical force. To determine this
relation, one needs to vary both the side slip angle as well as the vertical force. The side slip
angle has been varied between 1 and 11 with increments of 2 for a vertical load approximately
4000 N and 6000 N . Figure A.3 shows the measurements for a vertical load of 3983 N including
a fit based on a first order ODE for the tire relaxation


x
Fy = Fyss 1 e ,

(A.6)

with Fyss the steady state side force, x the distance the tire has traveled since the side slip angle
has been set and the relaxation length. The peaks occurring in the 11 are caused by the
tire completely sliding over the road surface. Plotting the steady state lateral force and relaxation
length as a function of side slip angle results in Figure A.4. For both vertical loads the typical peak
in side force can be observed at approximately 9 , furthermore, the relaxation length decreases
for increasing side slip.
77

Side force
5000

Fy [N ]

4000
3000
2000
1000
0

10

12

Relaxation length [m]

Relaxation length
0.5
3983 N
5890 N

0.4
0.3
0.2
0.1
0

10

12

Side slip angle [deg]


Figure A.4: Steady state lateral force and relaxation length as a function of side slip angle.

78

A.3 Magic Formula


The typical shape of the steady state side force as is shown in Figure A.4 can be approximated by
various models. Here, the Magic Formula tire model will be used as developed by Pacejka [43].

Fy = D sin C arctan B 0 E B 0 arctan B 0
(A.7)
The magic formula tire model contains various parameters that define the tire properties. For the
lateral force these parameters will be further explained in this chapter. First of all some general
parameters:
Adapted nominal load

Fz0
= Fz0 F z0

Nominal vertical load increment


d fz =

(A.8)

Fz Fz0

(A.9)

Fz0

All the parameters pi and j in the following equations are parameters that define the tires
properties and differ for each tire. Both the shift factors and camber will be omitted.

Shape factor C
The shape factor C defines the limit range of the sine function and thereby the shape of the total
function. Its quantity is calculated by:
C y = pC yl C y

(A.10)

Peak value D
The peak value D determines the height of the peak as is visible in Figure A.4. It is calculated by:
D y = y Fz

(A.11)

here, the friction coefficient y is defined as:



y = p Dy1 + p Dy2 d f z y

(A.12)

Curvature factor E
The curvature factor is:
E y = p E y1 + p E y2 d f z


1 p E y3 sign y E y

(A.13)

Stiffness factor B
The stiffness factor B is defined as:
By =
with:

Ky
C y Dy



K y = p K y1 Fz0 sin 2 arctan

79

Fz
p K y2 Fz0 F z0

(A.14)

F z0 K y

(A.15)

A.4 Tire parameters


Using the MF tire model as introduced above, a fit can be made using the Matlab routine fmincon.
This results in the tire parameters presented in Table A.1. Some parameters could not be determined due to the lack of measurement data including camber angles.
Table A.1: MF tire parameters, based on measurements from 07-09-2010. Parameters marked with an * were not determined because no camber angle was considered.

Parameter
C y1
y
v
E y
K y

pC y1
p Dy1
p Dy2
p Dy3
p E y1
p E y2
p E y3
p E y4
p K y1
p K y2
p K y3
V0
Fz0
kt y

Value
1
1
1
1
1
0
1.12
0.822
0.184
*
3.08 1011
7.43 1011
1.70
*
30.9
2.04
*
*
4000
271

Unit
rad
[-]
[-]
[-]
[-]
[-]
[-]
[-]
[-]
[-]
[-]
[-]
[m/s]
[N]
[kN/m]

80

Description

y shaping factor
Load dependent part of y
Camber dependent part of y

Nominal load
Lateral tire stiffness

Appendix B

LDIA 2011 digest


The following digest has been accepted. to the Eighth International Symposium on Linear Drives
for Industry Applications (LDIA 2011) which will be held on July 3-6 2011 in Eindhoven.

81

Robust control of a direct-drive electromagnetic active suspension system


T.P.J. van der Sande , B.L.J. Gysen , I.J.M. Besselink , J.J.H. Paulides, E.A. Lomonova and H. Nijmeijer
Eindhoven University of Technology, P.O. Box 513, Eindhoven, 5600MB,The Netherlands
email: t.p.j.v.d.sande@student.tue.nl
A BSTRACT
This paper considers the control of an electromagnetic active suspension system based upon a quarter
car model. Because variation of the sprung mass, tire
stiffness and damping exist and can not be estimated, a
robust control structure is considered. Depending on
the choice of the weighting filters, either comfort or
handling can be emphasized. Simulations as well as
measurements on a full size quarter car setup will be
considered, indicating the performance and efficiency
of the electromagnetic suspension system.
1

I NTRODUCTION

A conventional car suspension is always a trade-off between comfort and handling. Over the last decade, top of
the line manufacturers have therefore developed active suspension systems to enhance comfort when driving straight
whilst improving handling while cornering. Current day
examples are the ABC system [1] employed by Mercedes
which contains a hydraulic actuator in series with a passive suspension. Whilst this system can provide energy to
the suspension, a bandwidth of only 5 Hz is obtained and
continuous pressurization is required making the energy
demands very high. Another example is the Delphi magnetorheological damper [2] which, under the influence of
a magnetic field, can change its damping value within a
specified range. Benefits of this system are its high bandwidth and low power requirement. Energy can however
not be supplied to the system making this a semi-active
system. A solution to the drawbacks mentioned above is
using a tubular permanent magnet electromagnetic actuator [3] given its force density. Furthermore, it is capable
of delivering direct drive in a small volume and the bandwidth it can achieve is much higher than required to improve comfort and handling. Power consumption is lower
than that of a hydraulic system since no continuous pressurization is necessary and energy can even be recuperated
depending on the damping value and controller settings
[4].
Numerous publications exist for the control of active
suspension, for example Lee [5] considers lead-lag, LQ
and fuzzy control for a brushless tubular permanent magnet actuator. Due to the limited peak force (29.6 N) of
the actuator, a scaled down (sprung mass 2.3 kg, unsprung
mass 2.27 kg) test setup is considered making the setup
not representing a typical vehicle (sprung-unsprung ratio
10:1). Furthermore, the parameters of the setup are considered to be fixed and fully known. On the other hand,
Lauwerys [6] does use a full size quarter car setup and also
includes uncertainties in the design of the controller. However, with the actuator being hydraulic, only a reductions

Ft
Fact

Pertubed plant
White
Wi1 zr
noise
1

as

s/av+1

Fact

Wo1
LP
Wo2
HP

Weighted dynamic
tire load

Weighted actuator
force

Wo3
ISO2631

Weighted
sprung
acceleration

au
zs zu Wo4
noise

Wn1

noise

Wn2

Wn3

Weighted
suspension
travel
noise

Controller

Figure 1: Schematics used for design of robust control.

in the sprung resonance is considered (1.5 Hz) opposed to


the region where humans are most sensitive (4-8 Hz). Furthermore, RMS power requirements of the hydraulic suspension system are 500 W per corner, making the system
inefficient. To prove the increased efficiency, higher bandwidth and performance of an electromagnetic suspension
this paper considers the robust control on a full scale quarter car model including variations that occur in practice.
Section 2 discusses the control topology used in this
paper, section 3 treats preliminary results followed by the
conclusion in section 4.
2

C ONTROL

A quarter car model is generally accepted as a good


way of estimating comfort of a car by means of its sprung
acceleration (as ), it can furthermore be used to estimate
the influence of road disturbances on tire load variations
(Ft ) and suspension travel. It is therefore chosen to implement this model to develop controllers for the active
suspension system. The baseline vehicle for simulations
is a BMW 530i, given its minimum and maximum sprung
weight (ms ), one parametric uncertainty can already be
defined, see Table 1. The nature of the tire gives rise to
another parametric uncertainty since, due to changing vertical force its stiffness (kt ) varies. Finally, the tubular permanent magnet actuator has an electromagnetic damping
which has a regressive character. Therefore the damping
(ds ) also varies within a certain range as a function of velocity (vs vu ).
Given the parametric uncertainties, the most suitable
control method is robust control [7]. With this control
method an optimal controller is found, whilst being stable for all possible uncertainties. The inputs of the controller are the sprung acceleration (as ), unsprung acceleration (au ) and suspension travel (zs zu ) as Fig. 1 shows.

Table 1: Quarter car parameters.

S IMULATION R ESULTS

By changing the weighting filters introduced in the previous section either comfort (as ) or handling (Ft ) can be
emphasized. Given the constraints on suspension travel
and RMS actuator force (1 kN) a 72% improvement in
comfort can be achieved deteriorating handling by 145%.
On a straight road this is not of concern, however when
cornering, handling is important and a 24% improvement
in handling can be achieved by choosing another controller
as Fig. 2 shows. This figure also shows the passive BMW
and possible other suspension settings that can be achieved
by scaling the damper and spring of the passive BMW.
The difference with the active suspension is that the suspension characteristic is fixed and only one point can be
chosen whereas with active suspension the objective can
change constantly, i.e. the whole line can be used. Active strut off indicates the performance of the eddy current
damper in the active suspension in combination with the
passive spring. Measurements will be performed on a full
size quarter car test setup to verify the simulations.
4

C ONCLUSION

A robust controller is developed for an electromagnetic


suspension system based on a quarter car model. Parametric uncertainties are included in the quarter car model accounting for possible variations in the plant. Weighting
filters are chosen such that either maximal comfort or best
handling is achieved given the constraints of maximum

200
180
160
140
120
100
80
60
20

40

A
ctive
strut
off
P
assiv
e BM
W
H
andli
n g op
timal

(Ft controlled)/(Ft passive) [%]

220

Possible passive performance

al
im
pt

Controller options
240

o
rt

Also visible are the white noise input, filtered by a first order low pass filter, resulting in the road disturbance zr as
input to the disturbed plant. Wn1 , Wn2 and Wn3 represent filters that indicate the amount of measurement noise
on the sprung acceleration, unsprung acceleration and suspension travel . Futhermore, as Lauwerys [6] showed, frequency dependent weighting functions can be used to obtain required behavior of the system. It has for instance
been shown that humans are most sensitive to vertical vibrations in the range from 4-8 Hz [8]. The ISO2631-1
weighting criterion has been used to emphasize human
sensitivity in this region. When considering small side
slip angles, variations in vertical tire load (Ft ) only influence the lateral tire force at lower frequencies. A low-pass
filter is therefore used, penalizing low frequencies. The
constraints of the system are given by the maximum RMS
actuator force of 1000 N and maximum suspension travel
which is not allowed to surpass the suspension travel of
the passive BMW under similar conditions. To limit actuator force, a high-pass filter is used such that, beyond the
frequencies of interest actuator force is penalized.

260

fo
om

Value
352.5 - 525.27 kg
55.3 kg
30e3 N/m
3.1e5 - 3.7e5 N/m
900 - 1700 Ns/m

Name
Sprung mass
Unsprung mass
Suspension stiffness
Tire stiffness
Damping

Parameter
ms
mu
ks
kt
ds

60
80
100
(as controlled)/(as passive) [%]

120

Figure 2: Performance of robust controllers compared with passive


BMW.

suspension travel and actuator force. A 72% improvement


in comfort limited by suspension travel or a 24% improvement in handling which is determined by maximum RMS
actuator force can be achieved with the active suspension
compared to the passive BMW suspension. Measurements
will be included to verify the results achieved in simulations.
R EFERENCES
[1] Mercedes ABC, http://500sec.com/abc-active-bodycontrol. Online, 2010.
[2] Delphi
magneride
semi-active
suspension,
http://delphi.com. Online, 2010.
[3] B. Gysen, J. Paulides, J. Janssen, and E. Lomonova,
Active electromagnetic suspension system for improved vehicle dynamics, Vehicular Technology,
IEEE Transactions on, vol. 59, pp. 1156 1163, March
2010.
[4] B. Gysen, T. Van der Sande, J. Paulides, and
E. Lomonova, Efficiency of a regenerative directdrive electrmomagnetic active suspension, VPPC
conference, 2010.
[5] S. Lee and W. Kim, Active suspension control
with direct-drive tubular linear brushless permanentmagnet motor, IEEE transactions on control system
technology, vol. 18, pp. 859870, July 2010.
[6] C. Lauwerys, J. Swevers, and P. Sas, Robust linear
control of an active suspension on a quarter car testrig, Control engineering practive, vol. 13, pp. 577
586, 2005.
[7] S. Skogestad and I. Postlethwaite, Multivariable feedback control. Wiley, 2007.
[8] ISO, ISO2631-1:1997:Mechanical vibration and
shock - Evaluation of human exposure to whole-body
vibration, tech. rep., International Organization for
Standardization, Geneva - Switzerland, 1997.

Вам также может понравиться