Licenta Mecanica

Download as pdf or txt
Download as pdf or txt
You are on page 1of 98

A Fully Flexible Valve Actuation

System For Internal Combustion


Engines
by

Junfeng Zhao

B.A.Sc., Tsinghua University, 2007

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF


THE REQUIREMENTS FOR THE DEGREE OF

MASTER OF APPLIED SCIENCE

in

The College of Graduate Studies

(Mechanical Engineering)

THE UNIVERSITY OF BRITISH COLUMBIA


(Okanagan)
August 2009

c Junfeng Zhao 2009
Abstract

Air pollution, global warming, and rising gasoline prices have lead govern-
ments, environmental organizations, and consumers to pressure the auto-
motive industry to improve the fuel efficiency of cars.
Since alternative fuels such as hydrogen are still quite far from being
commercially viable, improving the existing internal combustion engine is
still an important priority. Traditional internal combustion engines use a
camshaft to control valve timing. Since the camshaft is rigidly linked to
the crankshaft, engineers can optimize the camshaft only for one particular
speed torque combination. All other engine operating points will suffer from
a suboptimal compromise of torque output, fuel efficiency, and emissions. In
an engine with a camless valve actuation system, valve events are controlled
independently of crankshaft rotation. As a result, fuel consumption and
emissions may be reduced by 15% ∼ 20% and torque output is enhanced in
a wide range of engine speeds.
The Fully Flexible Valve Actuation (FFVA) system is our approach to
construct a camless valve actuation system. Within the limits of the dynamic
bandwidth of the system, it allows for fully user definable valve trajectories
that can be adapted to any need of the combustion process. The system
is able to achieve 8mm valve lift in 3.4ms, which is suitable for an engine
operating at 6000RPM. The valve seating velocity is similar to conventional
valve trains that achieve 0.2m/s at high engine speeds and 0.05m/s at engine
idle conditions. Finally, the energy consumption measured in an experimen-
tal test bed matches the friction losses of conventional valve trains and it
can further be improved by using an optimized motor.
This thesis describes the progress that has been made towards designing
this technology. A design methodology is derived and important operation
features of the mechanism are explained. Modeling and simulation results
show significant advantages of the FFVA over previously designed electro-
magnetic engine valve drives.

ii
Table of Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Background and Motivation . . . . . . . . . . . . . . . . . . 2
1.2 Principles of Camless System . . . . . . . . . . . . . . . . . . 4
1.3 Current State of Research . . . . . . . . . . . . . . . . . . . . 7

2 FFVA System Overview . . . . . . . . . . . . . . . . . . . . . 12


2.1 Actuator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Principles of Brushless DC Motor . . . . . . . . . . . . . . . 14
2.3 Valve Control Unit . . . . . . . . . . . . . . . . . . . . . . . 15

iii
Table of Contents

3 Actuator Design with a Stock Motor . . . . . . . . . . . . . 18


3.1 Mechanical Optimization . . . . . . . . . . . . . . . . . . . . 18
3.2 Motor Selection . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3 Trajectory Generation . . . . . . . . . . . . . . . . . . . . . . 22
3.4 Design Example . . . . . . . . . . . . . . . . . . . . . . . . . 26

4 Motor Optimization . . . . . . . . . . . . . . . . . . . . . . . . 30
4.1 Stator Saturation Constraint . . . . . . . . . . . . . . . . . . 30
4.2 Demagnetization Constraint . . . . . . . . . . . . . . . . . . 32
4.3 Linearity Constraint . . . . . . . . . . . . . . . . . . . . . . . 35
4.4 Energy Constraint . . . . . . . . . . . . . . . . . . . . . . . . 36
4.5 Motor Design Example . . . . . . . . . . . . . . . . . . . . . 38

5 Finite Element Analysis of DC Motor . . . . . . . . . . . . . 42


5.1 Brushless DC Motor Topology . . . . . . . . . . . . . . . . . 42
5.2 Maxwell Equations . . . . . . . . . . . . . . . . . . . . . . . 43
5.3 Governing Equations in 2D Electromagnetic Field . . . . . . 44
5.4 Finite Element Simulation for BLDC Motor . . . . . . . . . 46
5.5 Evaluation of Optimized Motor . . . . . . . . . . . . . . . . 49

6 Control Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . 52
6.1 3 Phase Inverter . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.2 Position (or PD) Controller . . . . . . . . . . . . . . . . . . . 54
6.3 States Observer . . . . . . . . . . . . . . . . . . . . . . . . . 54

7 System Simulation and Experimental Validation . . . . . . 61


7.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . 61
7.2 Simulation of Transition Performance . . . . . . . . . . . . . 62

iv
Table of Contents

7.3 Simulation of the Robustness Towards Parameter Variations 64


7.4 Simulation of Disturbance Rejection . . . . . . . . . . . . . . 66
7.5 Simulation of Low Resolution Sensor with a Kalman Filter . 66
7.6 Experimental Results . . . . . . . . . . . . . . . . . . . . . . 67

8 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
8.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
8.2 Suggestions for Future Works . . . . . . . . . . . . . . . . . . 80

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

v
List of Tables

3.1 Motor Selection . . . . . . . . . . . . . . . . . . . . . . . . . . 27

5.1 Comparison of motor parameters . . . . . . . . . . . . . . . . 50

7.1 Motor Parameters . . . . . . . . . . . . . . . . . . . . . . . . 63


7.2 Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
7.3 Performance comparison . . . . . . . . . . . . . . . . . . . . . 68

vi
List of Figures

1.1 Powertrain evolution of EU standard . . . . . . . . . . . . . . 2


1.2 Key technologies . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Fuel economy solutions . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Schematic p-V diagrams . . . . . . . . . . . . . . . . . . . . . 9
1.5 Honda VTEC mechanical valve actuation . . . . . . . . . . . 10
1.6 Hydraulic valve actuation . . . . . . . . . . . . . . . . . . . . 10
1.7 Solenoid valve actuation . . . . . . . . . . . . . . . . . . . . . 11
1.8 EMVD system . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.1 System structure . . . . . . . . . . . . . . . . . . . . . . . . . 12


2.2 Real object and 3D model of QB02302 . . . . . . . . . . . . . 13
2.3 Linkage structure between motor and valve . . . . . . . . . . 16
2.4 2D geometry of BLDC motor . . . . . . . . . . . . . . . . . . 17

3.1 Trajectories . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.2 Jerk limited smooth trajectory . . . . . . . . . . . . . . . . . 23
3.3 Winding variation for QB02302 . . . . . . . . . . . . . . . . . 29

4.1 Demagnetization curve . . . . . . . . . . . . . . . . . . . . . . 33


4.2 Flux density due to current and magnet . . . . . . . . . . . . 34
4.3 Diagram of optimization procedure . . . . . . . . . . . . . . . 41

vii
List of Figures

5.1 QB02302 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.2 Optimized motor . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.3 Mesh result . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.4 Flux density distribution of the optimized motor . . . . . . . 51

6.1 Diagram of control units . . . . . . . . . . . . . . . . . . . . . 52


6.2 Cascaded control structure . . . . . . . . . . . . . . . . . . . 58
6.3 Current controller . . . . . . . . . . . . . . . . . . . . . . . . 59
6.4 Position controller . . . . . . . . . . . . . . . . . . . . . . . . 60

7.1 System hardware . . . . . . . . . . . . . . . . . . . . . . . . . 62


7.2 Simulation of displacement trajectories . . . . . . . . . . . . . 70
7.3 Simulation of energy losses of different trajectories . . . . . . 71
7.4 Transition simulation of the optimized motor . . . . . . . . . 72
7.5 Simulation of parameter variations . . . . . . . . . . . . . . . 73
7.6 Simulations of disturbance rejection . . . . . . . . . . . . . . 74
7.7 Simulation of transition response with a 7 bit sensor and var-
ious filtering strategies . . . . . . . . . . . . . . . . . . . . . . 75
7.8 Simulations of measurement error using different filtering strate-
gies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7.9 Experimental trajectory results using QB02302 . . . . . . . . 77
7.10 Experimental energy loss using QB02302 . . . . . . . . . . . . 78

viii
Acronyms

FFVA=Fully Flexible Valve Actuation


VVT=Variable Valve Timing
EIVC=Early Intake Valve Closing
LIVC=Late Intake Valve Closing
EMCV=Electromechanical Camless Valvetrain
EMVD=Electromechanical Valve Drive
NdFeB=Neodymium Iron Boron
FEM=Finite Element Method
FEA=Finite Element Analysis
LPF=Low Pass Filter
PWM=Pulse Width Modulation

ix
Nomenclature

r Rotor Radius
hs Slot Height
hrs Thickness of Stator Back
rsi Stator Inner Radius
rso Stator Outer Radius
l Motor Length
lg Airgap
lge Effective Airgap
lm Magnet Thickness
β Magnet Coverage Span
wt Tooth Width
ws Slot Width
KT Motor Torque Constant
KB Back EMF Constant
R Motor Terminal Resistance
L Motor Terminal Inductance
Z Total Number of Conductors
I Current in each Conductor
Itotal Total Current Itotal = ZI
T Motor Torque
ra Length of Excenter Arm
Jm Motor Inertia
m Valve Mass
J Total Inertia J = Jm + mra2
α Angular Acceleration
ω Angular Velocity
θ Rotation Angle
a Linear Acceleration
v Linear Velocity
s Linear Displacement
j Jerk(Derivative of a )

x
Nomenclature

E Energy Loss
ρ Density of Core and Magnet
Br Magnet Residual Flux Density
Bc Flux Density in Core
Bs Saturation Flux Density
Bgo Airgap Flux Density Due to Magnet Alone
BI Airgap Flux Density Due to Current Alone
Bg Airgap Flux Density Bg =Bgo + BI
BD Demagnetization Flux Density
wp Width of Each Pole
wm Width of Each Magnet
Ks Linear Current Density
γ Angular Displacement Between the Fields Produced by the Magnet and
the Stator Current
kw Stator Length Factor
Aca Copper Area per Slot
Aslot Slot Area
dw Wire Diameter
nw Number of Wires per Slot
kf ill Filling Factor
p Number of Poles
q Number of Slot per Pole per Phase
Q Total Number of Slot


H Vector of Magnetizing Field


D Vector of Electric Displacement


J Vector of Free Current Density


B Vector of Magnetic Field


A Potential Vector
ε Electric Permeability
µ Magnetic Permeability
σ Electric Conductivity

−n Normal Unit Vector
Pi Proportional Gain in PI Controller
Ki Integral Gain in PI Controller
Pd Proportional Gain in PD Controller
Kd Derivative Gain in PD Controller

xi
Acknowledgements

I would like to thank all people who have helped and inspired me during my
master study.
I especially want to thank my advisor, Prof. Rudolf Seethaler, for his
guidance during my research and study at University of British Columbia
Okanagan. His perpetual energy and enthusiasm in research had motivated
me. In addition, he was always accessible and willing to help me with my
research. As a result, research life became smooth and rewarding for me. I
am indebted to him more than he knows.
Prof. Holzman, Prof. Najjaran, and Prof. Koch deserve a special thanks
as my thesis committee members and advisors.
Finally, I thank my family for supporting me throughout all my studies
at University of British Columbia Okanagan.

xii
Chapter 1

Introduction

In conventional Internal Combustion (IC) engines, the timing of intake and


exhaust valves is controlled by the shape and phase angle of cams. Engineers
need to choose the best compromise timing among fuel economy, emissions,
and torque, to design the shape of the cam. The optimization of cam shape is
possible only at one engine speed. But IC engines in automobile application
operate over speed and load ranges covering about an order of magnitude
in each variable. This wide range of speeds and loads results in conflicting
demands for the design of the lift profiles for the valves.
A fully flexible camless valve mechanism [1][2][3] allows controlling the
engine load without a separate throttle and thereby avoids the associated
energy loss, which will be discussed in next section. It is also known that
in a conventional engine the control of valve overlap, during which both the
intake and exhaust valves are open, can affect the emissions, full load and
idle performance. For example, to achieve high efficiency at high speed and
high load, a large amount of overlap is desired, however, this will not allow
the engine to idle smoothly at low speed and low load because the residual
fraction is excessive. Therefore, engine designers are starting to consider
camless systems. The Fully Flexible Valve Actuation (FFVA) system in our
research is a very promising technology of its kind.
Fully Flexible refers to the ability to control the duration (for how long
the valve is kept open or closed), the phase (when the valve should be opened
or closed), and the lift (how far does the valve move). Many presented
systems have only provided duration and phase control and they are referred
to as Variable Valve Timing (VVT) systems [4][5].
Flexible intake and exhaust valve mechanisms can greatly improve fuel
economy, emissions, and torque of the internal combustion engine. Fuel
consumption may be reduced by 15% − 20% [6], torque output is enhanced
in wide range of engine speed, and emissions may be decreased by the same
ratio.

1
1.1. Background and Motivation

Figure 1.1: Powertrain evolution of EU standard [1]

1.1 Background and Motivation

The emissions of automobiles are becoming a global problem. Many coun-


tries and areas have been focusing on them and have set different standards
to regulate emissions and to improve the performance of engines. Emission
standards are requirements that set specific limits to the amount of pollu-
tants that are released into the environment. While emission performance
standards have been used to dictate limits for conventional pollutants such
as oxides of nitrogen N Ox , this regulatory technique may also be used to
regulate fuel consumption which is related to the emission of carbon dioxide
CO2 .
Emissions are directly influenced by the combustion process. A fuel that
burns incompletely is obviously not being utilized efficiently, and unburnt
fuel remains as pollutants. In both gasoline and diesel engines, the fuel is
atomized very finely to ensure good combustion. But that’s just one part
of the story. The other is that the nature of the combustion process also
influences pollutant emissions.
To address the problem of global warming, emissions of exhaust pollu-

2
1.1. Background and Motivation

Figure 1.2: Key technologies [1]

tants and greenhouse gas from motor vehicles must be reduced [7]. Fig-
ure 1.1 shows the EU standard of powertrain evolution in the future for
both gasoline engines and diesel engines, which are represented by red and
black spots respectively in the picture. These objectives are challenging and
conflict with each other. It can be seen that for both gasoline and diesel
engines there is still a long way to go before they can meet the proposed
standard. In practice, if N Ox emissions are reduced, the CO2 emissions
and fuel consumption inevitably rises. With the more stringent regulations
expected, new engine technologies are urgently required.
In Figure 1.2, key technologies for improving engine’s performance are
shown. The circled camless system, is on a very promising place on the map.
Also, it can be seen from Figure 1.3 that the camless technology is the most
effective, since it can improve the fuel economy by 15% ∼ 20%.
But a combination of different technologies needs to be developed, since
no single one can fully satisfy the future standard. The large number of mu-
tually interactive variables and sub-systems is a major engineering challenge,
but it also provides the flexibility that is essential for converging to an op-
timal solution. After all, a properly matched combination of fuel, injection
system, air management, and exhaust treatment should make it possible to

3
1.2. Principles of Camless System

Figure 1.3: Fuel economy solutions [1]

create sufficiently stable operating conditions to meet prevailing regulatory


emission and fuel economy targets.
Another advantage of camless technology is that it is compatible with
all other energy saving solutions. Thus, the research on camless systems is
necessary and important for the evolution of engines.

1.2 Principles of Camless System

In a conventional four stroke internal combustion engine, the operating cycle


takes place over two revolutions of the crankshaft [9].
In the first stroke, the piston of engine starts from top dead center
(TDC). At the beginning of the movement, the intake valve is already
opened, which is called inlet valve opening (IVO), to let the fresh air or
air-fuel mixture into the cylinder. Then after the piston has reached the
bottom dead center (BDC), the intake valve is closed, which is called inlet
valve closing (IVC).
The second stroke from BDC to TDC is defined as a compression stroke,

4
1.2. Principles of Camless System

during which both of the intake and exhaust valves are closed. Then the
mixture in the cylinder is compressed when the piston moves upwards. Ig-
nition is started when the piston is close to the end of the stroke. Then the
pressure in cylinder increases rapidly.
The third stroke from TDC to BDC is defined as a power stroke. When
combustion is completed, the pressure in the cylinder reaches a maximum
value. Then the cylinder starts expanding, which pushes the piston to move
until the exhaust valve opens, which is called exhaust valve opening (EVO).
Generally before BDC, the first portion of the burned gases is allowed to be
discharged.
The fourth stroke from BDC to TDC is defined as an exhaust stroke, as
the piston pushes the remaining burned gases out. Near the TDC position,
the intake valve opens again and the exhaust valve closes, which is called
exhaust valve closing (EVC). The timing is usually slightly after the TDC
position.
Figure 1.4 shows schematic p-V diagrams of a four-stroke engine with
a conventional mechanical valvetrain (plot 1) and a camless systems with
early intake valve closing (EIVC, plot 2) and late intake valve closing (LIVC,
plot 3). The shadowed area corresponds to the work needed to induct and
expel gases. These losses are referred to as pumping losses.
The pumping losses due to the throttling operation in a conventional
engine with a mechanical valvetrain are quite significant [10]. A challenge in
improving engine efficiency is to reduce the pumping losses while the friction
loss is not excessively increased. Two strategies are available for controlling
the amount of air drawn into the combustion chamber with variable valve
timing systems: Early intake valve closing (EIVC) and late intake valve
closing (LIVC). With EIVC, the engine load is controlled by closing the
intake valve early to trap the desired amount of charge, instead of throttling
the incoming charge conventionally by means of a throttle plate. Likewise,
LIVC also controls engine load with reduced pumping loss, but in this case
by returning unwanted charge to the intake manifold.
The comparison between pressure cycles clearly shows the advantages
obtainable by adopting a camless valve actuation system in the same engine
due to a significant reduction in pumping losses.
No matter whether EIVC or LIVC is implemented in the combustion
process, the full controllability of each valve must be obtained by replacing
the conventional valve system with a camless system, which must have some

5
1.2. Principles of Camless System

equivalent performance, while allowing the additional flexibility mentioned


previously.
First, a camless system must allow for fast valve transitions, where the
transition time refers to the time required to either open or close a valve.
At an engine speed of 6000rpm, the nominal cycle time for a conventional
four stroke engine is 20ms/cycle. Each cycle includes a closed-to-open tran-
sition, an open-to-closed transition, and holding phases between the transi-
tions. The valve should be open for about one third of the cycle. Thus, the
transition time at 6000rpm engine speed needs to be about 3.5ms [11].
The second constraint is the valve seating velocity which is the valve
speed when the valve hits the cylinder head after the valve closing transition.
In a typical IC engine, the seating velocity is usually less than 0.3m/s [12] at
high speed and less than 0.05m/s [13] at idle. Higher valve seating velocities
lead to excessive noise and potentially damage to the engine. Thus, ”soft
landing” is an essential requirement for any valve actuation system.
The third consideration is energy loss for each transition. In a typical
2.0L, 16 valve, four stroke IC engine, the total energy loss of the valve train
at 6000rpm is about 2−3kW [14]. If only considering the eight intake valves
(due to gas force, there is additional energy loss for each exhaust valve, but
we consider intake valves only here), the power loss for each intake valve is
around 1kW . Thus, at 6000rpm, and 20ms/cycle, the energy loss for each
intake valve is approximately 2.5J/cycle. Therefore, the average energy loss
of each valve at 6000rpm is approximately 1.25J/transition. This assumes
that the power to keep the valve open or closed is minimal, which is the case
for all valve actuation systems proposed to date.
The three constraints above are the main indicators of the system per-
formance. Additional constraints (such as working temperature, applied
voltage and system package size etc.) should be carefully considered before
a system can be put into practical application.
Different vehicle engines run different temperatures at different condi-
tions, but a fairly normal range is around 80o C to 140o C [15]. So the
practical system should be able to work at this temperature.
Several voltage standards [16] currently exist for automotive applica-
tions: most automobiles use 12V and trucks often use 24V . In order to
reduce the size of the wire harnesses and reduce copper losses, a 42V stan-
dard has been proposed but not implemented in production. In addition,
Hybrid cars usually can provide approximately 200V onboard voltage. For

6
1.3. Current State of Research

our investigation we will focus on 42V .

1.3 Current State of Research

In recent years, some relatively basic variable valve actuation systems have
been presented in publications or even introduced into production engines.
These forms mainly use mechanical, electro-hydraulic or electromagnetic
valve actuation technologies.
The mechanical systems could be very simple in structure as in Fig-
ure 1.5. The simplest ones, the cam phasers, change only valve timing.
Other types, slightly more complex, can change the valve lift. These mech-
anisms are quite simple and effective, and some of them are already on the
market. But the mechanical mechanisms can only provide limited flexibility
of the valve’s motion, and their dynamic response is too slow for guaran-
teeing optimum valve timing for transient engine operations. The Honda
VTEC mechanism [17] and the Toyota VVT-i system [18] are examples of
mechanical variable valve timing systems.
Electro-hydraulic systems [21] [22] are conceptually quite simple. The
electro-hydraulic camless systems proposed so far usually offer a contin-
uously variable and independent control of all aspects of valve motion.
Electro-hydraulic systems typically use piezo-actuated valves to control the
hydraulic fluid flow that is used to displace the valve. Unfortunately, hy-
draulic systems suffer from viscosity changes across the required temperature
range, since engine oils are typically used as the hydraulic liquid. Thus, the
performance deteriorates at low temperatures. In addition, it is very dif-
ficult to achieve good energy efficiency with hydraulic systems, since there
is no simple way to recover the kinetic energy of the valves when they are
slowed down [23].
Electromagnetic systems are characterized by a spring system that is
used to accelerate and decelerate the valve. Magnets or motors are used to
hold the valves in the end positions and to compensate for friction loss as
well as for combustion forces. Devices using solenoids such as in the one
shown Figure 1.7, are able to generate flexible valve phase and duration,
however, high valve seating velocity is difficult to control and the valve lift
is limited by the structure. Nevertheless, this is a very popular technology
and needs to be discussed in more detail.

7
1.3. Current State of Research

Typically, the solenoid actuator consists of a linear-moving armature


with two coils and two preloaded springs. The springs can achieve rapid
transition times while minimizing electrical energy input and are essential
in overcoming the significant combustion pressures. The electromagnets are
required for ”catching” the armature at either stroke bound. In addition,
they are used to overcome friction and pressure disturbances. This device
needs only very little current to hold the valve at both ends. But since there
is a very nonlinear relationship between force, position and current, it is
very difficult to control the seating velocity when the armature approaches
either end [6].
Figure 1.8 shows another electromagnetic system, in which a BLDC mo-
tor is applied to drive the valve rather than solenoids. A number of different
configurations [14][25][26] exist for this design that all use springs to accel-
erate and decelerate the valve and a motor driven pivoting cam to provide
timing. Note that the cam has a constant radius at either end of the valve
motion. As a result, the motor can keep the valve at either end using
zero torque. The disadvantage of this system clearly is its relatively high
mechanical complexity, and the inability to adjust valve lift continuously.
Nevertheless, the EMVD system demonstrates good performance on both
transient time and seating velocity.
Since electromechanical systems provide better energy recovery poten-
tial, they will be used as benchmarks for the proposed FFVA system. It
combines advantages of the other electromechanical systems and avoids some
of their inherent problems. Like EMVD, FFVA uses a BLDC motor to gen-
erate shear force to drive the valve. This leads to a much simpler linear
control system than that for the EMCV system. In contrary to EMCV and
EMVD, FFVA does not use springs, but energy recovery is provided by the
motor that is able to electrically feed back the breaking energy to storage
capacitors. In the following sections, the system structure, design procedure
and system performance are discussed in detail.

8
1.3. Current State of Research

Figure 1.4: Schematic p-V diagrams [8]

9
1.3. Current State of Research

Figure 1.5: Honda VTEC mechanical valve actuation [19]

Figure 1.6: Hydraulic valve actuation [20]

10
1.3. Current State of Research

Figure 1.7: Solenoid valve actuation [24]

Figure 1.8: EMVD system [14]

11
Chapter 2

FFVA System Overview

A schematic of the FFVA system is shown in Figure 2.1. There are three
main parts: the actuation part, the valve control unit and the engine control
unit.

Figure 2.1: System structure

The engine control unit controls the engine operation and provides the
required valve timing information to the valve control unit. The engine
control unit is designed separately, and not part of this research.

12
2.1. Actuator

Figure 2.2: Real object and 3D model of QB02302

The valve control unit receives the desired valve timing and lift from the
engine control unit and uses this information together with measured valve
position and measured current in order to regulate the amount of voltage
applied to the actuator.
The actuator consists of BLDC motor, a valve and a linkage structure
connecting the two. A state observer, including sensors and digital filters,
is connected to the actuator and used to measure the valve motion, which
is then provided to the valve control unit.

2.1 Actuator

Figure 2.3 shows a detailed picture of the linkage between motor and valve.
An excenter arm of length, ra , is attached to the motor’s shaft, which in turn
is connected to the valve through a small bracket. This structure transfers
the motor’s angular movement to the valve’s linear movement:

a = αra (2.1)

If the rotor angle is small, the relationship between lift and rotor angle
can be assumed linear. The excenter arm is made of aluminium and its
inertia can be neglected when compared to the rotor inertia and the valve
mass. The motor torque T is used to accelerate the rotor with an inertia,
Jm , and the valve with a reflected inertia mra2 , where m is the mass of the
valve:
T = (Jm + mra2 )α (2.2)

13
2.2. Principles of Brushless DC Motor

The figure also shows several parameters, which play an important role
in determining the system’s performance. These parameters, especially the
rotor angle θ and excenter arm length ra , will be discussed in the chapter of
optimization.

2.2 Principles of Brushless DC Motor

As shown above, a brushless DC motor is the key part of the FFVA system.
The basic structure of a permanent magnet brushless DC motors has three
elements: a stator with windings, a rotor with permanent magnets attached
to it, and a sensor to measure the rotor position. A picture of QB02302 [27]
and its 3D model is given in Figure 2.2.
The motor’s torque constant and other properties are all related to the
structure and material of the motor. Figure 2.4 provides the dimensioning
details of an interior rotor BLDC motor (2D geometry of motor QB02302).
The rotor radius r, the slot height hs and the length of the motor l are treated
as variables for an optimized design goal in subsequent chapters. The airgap
lg and magnet thickness lm are selected to give sufficient flux density in the
airgap. Magnets are glued on the rotor surface. The coverage of the magnets
on the rotor is called span, represented by β. In our application, the span is
about 0.8. Inside the stator, the tooth width wt and the slot width ws are
almost the same.
To understand how these elements work as a motor, consider some ele-
mentary magnetics [28]. When a current carrying wire is placed in a mag-
netic field so that the current flow is perpendicular to the direction of the
field, a force is exerted between the field producing element and the wire.
The electromagnetic model of the actuator is approximated by an equiv-
alent linear single phase DC-motor model. The mathematical model of the
BLDC motor is based on the following assumptions: 1) stator resistances
of all the three phases are equal and the self and mutual inductances are
constant; 2) the motor is operated within the rated condition and hence the
saturation effect due to current level is neglected; 3) iron losses are negligible.
This approach is followed, since most motor specifications are based on
this model and we want to be able to select an optimum motor from the
manufacturer’s motor specifications. The motor current I, is defined by
a first order differential equation in terms of the applied voltage U , and

14
2.3. Valve Control Unit

the back EMF voltage KB ω, the winding resistance R, and the winding
inductance L:
dI 1
= (U − KB ω − IR) (2.3)
dt L

The acceleration capability of the device is the torque T , which is pro-


portional to the current I:
T = KT I (2.4)

The maximum or peak torque available for a particular motor is given


by:
Tmax = KT Imax (2.5)

This is not the continuous torque available, which is usually constrained


by heat generation of the windings, but the value of torque that is con-
strained by saturation of the core or demagnetization of the permanent
magnets on the rotor. It is important to recognize that actuating valves
is an intermittent operation in nature and requires large accelerations over
short periods of time. Thus, the performance is governed by the peak torque
and resistive copper losses of the motor rather than the continuous torque
rating of the motor. For small automotive engines, the actuator motor
should consume less than 120W at 6000rpm and provide enough transient
torque in order to achieve valve openings or closings in 3.5ms. Chapter 3
will show how to select a suitable stock motor for this application. Chapter
4 provides a strategy to modify the stock motor in order to achieve optimum
performance.

2.3 Valve Control Unit

The main function of the valve control unit is to move the valve from the
closed to the opened position (and vice-versa) avoiding noise, which is caused
by nonzero seating velocity. This is achieved using a cascaded tracking
controller with feed forward.
In Chapter 3, valve trajectories for minimizing energy and maximizing
acceleration will be derived, and in Chapter 6 the design of the tracking
controller will be described.

15
2.3. Valve Control Unit

Figure 2.3: Linkage structure between motor and valve

16
2.3. Valve Control Unit

Figure 2.4: 2D geometry of BLDC motor

17
Chapter 3

Actuator Design with a


Stock Motor

The FFVA system is a servo system that uses a lightweight mechanical


linkage structure to transfer the motor rotation to the valve translation.
Typically, high speed servo systems operate at the accelerations less than
40g. The application shown here requires acceleration in the order of 250g.
At the same time, energy consumption needs to be minimized in order to
ensure that fuel consumption reductions gained through the introduction of
variable valve control are not offset by the power consumption of the valve
actuation system. To achieve these goals, the mechanical linkage, the electric
motor, and the valve trajectories are optimized in the following sections.

3.1 Mechanical Optimization

In the FFVA system, the motor’s rotation is converted to valve’s vertical


motion through an arm-like structure called the excenter arm. The length
of the excenter arm, ra , plays an important role in the performance of the
actuation system. This section outlines how the arm length can be optimized
to provide maximum acceleration and minimum energy consumption.
Maximum power transfer will occur in a mechanical system if the inertia
of the load matches the inertia of the motor. That is, for a specific motor,
if load inertia reflected to the motor shaft can be made to match the motor
inertia, disregarding added inertia and inefficiency of the reducer, power
transfer will be optimized and maximum acceleration of the load will result.
The proof for this is detailed [29].
The motor torque, T , is used to accelerate the rotor with inertia, Jm ,
and the valve with a reflected inertia of mra2 , where m is the mass of the

18
3.1. Mechanical Optimization

valve:
T = (Jm + mra2 )α (3.1)

This equation can be rearranged to provide the rotor acceleration, α:


T
α= (3.2)
Jm + mra2

The valve acceleration is a = α · ra . Taking the derivative of a with


respect to ra results in:
da d T ra T (Jm + mra2 ) − 2T mra2
= ( ) = (3.3)
dra dt Jm + mra2 (Jm + mra2 )2

By setting the derivative equal to zero, one finds the excenter arm length
that provides maximum acceleration:
r
Jm
ra = (3.4)
m

It will now be shown that the same excenter arm length also provides
minimum energy consumption, given a desired valve acceleration profile.
When the motor’s rotation angle, denoted as θ, is small enough, the valve
motion can be regarded as linear. Then the relationship between valve’s
linear displacement s and motor’s rotational displacement θ can be expressed
as:
s = ra θ (3.5)
Thus,
d2 s
= ra · α (3.6)
dt2
where α is angular acceleration. The torque that the motor provides is:
J d2 s
T = Jα = (3.7)
ra dt2
where J is the total inertia including motor inertia Jm and load inertia mra2 :

J = Jm + mra2 (3.8)

If the motor constant is KT , the current required is:


T Jα J d2 s
I= = = (3.9)
KT KT KT ra dt2

19
3.2. Motor Selection

Now consider the motor’s copper loss, which depends on current in the
form of:
d2 s 2 Jm + mra2 2 d2 s
Z Z Z
2 J 2
E = I Rdt = ( ) R ( 2 ) dt = ( ) R ( 2 )2 dt
K T ra dt K T ra dt
(3.10)
Take the derivatives on both sides:
Jm Jm + mra2 2 d2 s 2
Z
∂E 2R
= 2 (m − 2 )( ) R ( ) dt (3.11)
∂ra KT ra K T ra dt2

Then the optimal arm length is found as:


r
Jm
ra = (3.12)
m
which is exactly the same as the optimal arm length for maximum acceler-
ation derived previously.
This suggests that matching the inertia of the load with the inertia of
the motor will simultaneously provide maximum acceleration and minimum
energy consumption.

3.2 Motor Selection

Few off-the-shelf motors are designed for the highly dynamic operating con-
ditions found in the FFVA system. However, it is very important for a suc-
cessful implementation that the motor operates with high efficiency. Thus,
using the findings from the mechanical optimizations, criteria for selecting
a motor that provides maximum valve acceleration and minimum energy
consumption are derived in the following section.
The mechanical optimization assumed that the rotor motion would be
small enough in order to ensure a linear relationship between valve and
motor motion. Using this assumption, the identical Equations 3.4 and 3.12
describe the optimal excenter arm length. One can now go backwards and
describe the minimum motor inertia required in order to achieve a rotor
angle smaller than θmax , for a maximum valve motion, smax , and a valve
mass, m:
smax 2
Jm > m( ) (3.13)
θmax

20
3.2. Motor Selection

Since minimizing the size of the motor reduces cost and facilitates pack-
aging the motor in the cylinder head, one would usually choose motors with
inertia close to the linearity constraint.
In addition to the linearity requirement, the motor is also required to
provide large acceleration. Substituting Equation 3.4 back into Equation 3.2
provides an equation for the maximum valve acceleration in terms of motor
and valve parameters:
Tmax ra Tmax
amax = = √ (3.14)
Jm + mra2 2 Jm m

This expression indicates that maximum acceleration is achieved by min-


imizing the valve mass. It also shows that the ratio of maximum torque over
square root of motor inertia should be maximized. The relationship can be
used to compare and select stock motors from their specification data sheets.
In addition to large accelerations, the motor also needs to minimize cop-
per losses.
If the optimal arm length is substituted back into Equation 3.10 for
energy consumption, the following equation is obtained:

d2 s 2
Z
Jm R
E = 4m × × ( ) dt (3.15)
KT2 dt2

This equation indicates that there are three parts of the actuator that
need to be optimized in order to minimize copper losses.

• Energy consumption is directly related to valve mass m. A low mass


valve will significantly reduce energy consumption.

• The energy cost term of the motor consists of three parameters, the
motor inertia multiplied by the electrical resistance divided by the
square of the motor torque constant. The energy cost term provides
a convenient guideline to compare and select stock motors using their
specification data sheets. Chapter 4 will derive an optimum design
procedure for the motor.

• The last term in Equation 3.15 is the integral of the acceleration tra-
jectory of the valve. The next section will show how to design energy
optimal trajectories.

21
3.3. Trajectory Generation

3.3 Trajectory Generation

The usual requirement for a valve trajectory is that the valve needs to travel
the desired lift within a prescribed transition time t4 , which is given by the
engine control unit. Both lift, and transition time will vary with changing
engine speeds and engine torque requirements. Thus, a simple online al-
gorithm that generates energy optimal valve trajectories is required. The
problem can be solved by calculus of variations [30], which leads to the
following valve acceleration [31]:
2t
a(t) = amax (1 − ) (3.16)
t4
where amax is the maximum acceleration required to achieve the desired

Figure 3.1: Trajectories

valve lift and transition time with minimum energy consumption. The wave
form is shown as the plot of ”Optimal” in Figure 3.1 and the resistive copper

22
3.3. Trajectory Generation

Figure 3.2: Jerk limited smooth trajectory

losses for this trajectory can be expressed as:

3s2 (Jm + mr2 )2 R


Z
Eopt = I 2 Rdt = max 3 2 2 a (3.17)
2t2 ra KT

where smax is the maximum valve travel, and t2 = t4 /2 is half the valve
travel time.
In practice, the optimal trajectory suggested above is not feasible due to
two physical constraints.
First, maximum current constrains the available acceleration for a motor

23
3.3. Trajectory Generation

with constant KT [32]:


T ra IKT ra
a= = (3.18)
J J
Secondly, the ideal trajectory contains acceleration discontinuities, or
infinite jerk at the beginning and at the end of the valve movement. However,
jerk is limited by the driving voltage:
K T ra v aJ
j= (Umax − KT −R ) (3.19)
JL ra K T ra

If the motor parameters are fixed, then the voltage becomes the limiting
factor for jerk. The optimal trajectory can be followed only when the motor
can provide infinite jerk, which requires infinite voltage.
The remaining section then outlines a procedure for imposing a jerk lim-
ited smooth trajectory. The kinematic profiles used in trajectory generation
are illustrated in Figure 3.2. The suboptimal acceleration has a triangular
profile with maximum slopes (i.e. jerk values) at the start and at the end
of the valve motion. It requires a slightly higher amax than the optimal
trajectory, which can be seen from Figure 3.1.
Given a time t1 during which the initial maximum jerk j1 is applied, the
value of this maximum jerk is defined as:
3smax
j1 = t 1 t2 (3.20)
2t2 − t1

The energy lost in the copper windings during a single transition from
closed to open or from open to closed, with the sub-optimal trajectory is
given by:
6s2max J 2 R
Z
Esubopt = I 2 Rdt = (3.21)
2t2 (t1 − t2 )2 ra2 KT2

It is interesting to note the ratio of energies of the optimal to the sub-


optimal trajectory. As expected the ratio increases for larger values of t1 ,
which leads to the conclusion that a motor with rather small inductance is
required in order to achieve good energy efficiency:
Eopt 4t22
= (3.22)
Esubopt (t1 − 2t2 )2

We now aim our attention towards finding the minimum value for t1 .
Equation 3.19 indicates that during the initial period where j = j1 , the

24
3.3. Trajectory Generation

voltage must be continuously increasing, since the jerk is constant and both
velocity and acceleration are increasing. Thereafter, the voltage decreases,
since jerk is reversed. Thus, it is expected that the maximum voltage will
take place at t = t1 . Equations 3.19 and 3.20 for t = t1 lead to a quadratic
equation in terms of t1 :

(3KT2 smax +2KT t2 ra U )t21 +6smax LJ +(6smax RJ −4KT t22 ra U )t1 = 0 (3.23)

The solution to this equation represents the minimum and maximum


value for t1 . The smaller of the two values is the minimum energy solution
for the triangular trajectory. Note that the fastest trajectory possible with
a given motor and fixed supply voltage is the one where the solution to
Equation 3.22 reduces to a single real value. The solution to this problem
is a quartic equation in terms of t2 that is difficult to solve analytically. In
practice however, one usually is presented with the problem of finding j1 and
t1 for a desired valve travel time. Having picked the largest possible value of
j1 , one would usually check whether the required acceleration at time t = t1
does not violate the torque constraint. If the required acceleration is too
high, the transition time is too short or the valve lift is too high.
Given that the initial displacement, velocity and acceleration are zero,
and values for j1 and t1 from Equations 3.20 and 3.23, the acceleration a,
velocity v, and displacement s profiles can be obtained by integrating the
jerk profile:
Z t
a(t) = a(ti ) + j(t)dt;
ti
Z t
v(t) = a(ti ) + a(t)dt;
ti
Z t
s(t) = v(ti ) + v(t)dt
ti

The jerk profile in Figure 3.2 can be written as:



j1
 0 ≤ t < t1
j(τ ) = j2 t1 ≤ t < t 3 (3.24)

j1 t3 ≤ t < t 4

where j1 , j2 are the magnitudes of jerk in different regions. Integrating


Equation 3.25 with respect to time, and denoting the maximum acceleration

25
3.4. Design Example

and deceleration magnitudes with a1 and a2 respectively, the acceleration


profile can be expressed as:



 j1 (t − t1 ) 0 ≤ t < t1

a + j (t − t ) t ≤ t < t
1 2 2 1 2
a(t) = (3.25)


 j2 (t − t3 ) t2 ≤ t < t 3

a + j (t − t ) t ≤ t < t
2 1 4 3 4

The same idea can be applied to the velocity and position profiles. For
constant jerk, acceleration profiles are linear, velocity profiles are parabolic
and displacement profiles are cubic. The advantage of using a jerk limited
profile is that triangular acceleration profiled trajectories have smoother
velocity, acceleration and jerk characteristics compared to other profiled
trajectories. The control signals resulting from the utilization of such a ref-
erence trajectory will also be smoother, hence reducing the risk of impacting
the drives, or exciting the machine’s structural dynamics.
In the simulation of FFVA system, a Waveform Generator is programmed
to generate the reference trajectory. The transition timing and the displace-
ment of the valve can be set.

3.4 Design Example

From the sections above, two important criteria for motor selection were
derived which are summarized as follows:

Tmax
• amax = √
2 Jm m

Jm R
• E∝ KT2

The first criteria shows the maximum acceleration is constraint by two


motor parameters: The maximum torque, and the inertia of the motor. The
energy consumption for a fixed trajectory and a fixed valve mass provides
the second rule to guide the motor selection process.
In addition, energy consumption for a desired transition time with a
suboptimal triangular trajectory can also be compared using Equation 3.21.

26
3.4. Design Example

Comparison with other proposed models


Motor Name QB02302 119007 I2383092NC
Manufacturer Maccon[27] Pittman[33] MCG[34]
Torque Constant KT [Nm/A] 0.076 0.0718 0.066
Terminal Resistance R [Ω] 0.24 0.541 1.12
Motor Inertia Jm [kg.m2 ] 2.3E-5 3.1E-5 1.7E-5
Peak Torque Tmax [Nm] 7.6 2.19 1.235
T
Acceleration Term √Jmax
1585 393 299
m
Jm R
Energy Cost Term K 2 9.56E-6 32.5E-6 43.7E-6
T
Energy Cost E 2.68 9.12 12.3
for 6000rpm [J/Cycle]

Table 3.1: Motor Selection

In Table 3.1, three candidates with similar size are listed. The size is
chosen in order to provide the minimum inertia that fulfills the linearity con-
straint between valve movement and motor movement. Since the QB02302
series motor outperforms the other candidates in terms of energy consump-
tion and acceleration, it is used in the experimental test bed in Chapter
7.
It should be pointed out that the QB02302 motor can be ordered with
different windings.
To make a first approximation, changing the windings from N to N ∗
affects the motor parameters as follows:
N∗ N∗ 2 N∗ 2
KT∗ = KT , R ∗ = ( ) R, L∗ = ( ) L (3.26)
N N N

Note that the energy cost term is not affected by changing the number
of windings. Also, the maximum torque and thus, the available acceleration
does not change, even though the current required to achieve this torque will
be altered, since the torque constant KT is linearly related to the number
of windings:
N
Imax = Tmax (3.27)
KT N ∗

27
3.4. Design Example

Another interesting relationship for the FFVA system is the jerk available
for a given maximum supply voltage Umax :
K T ra N KT v aJ
j= (Umax ∗ − −R ) (3.28)
JL N ra K T ra

Equation 3.28 indicates that the available jerk scales linearly with volt-
age and the inverse of the winding ratio. In addition, the jerk is reduced
once the valve starts moving. To summarize, decreasing the number of wind-
ings increases the available maximum jerk. However, the current required to
achieve the maximum acceleration is increased at the same time. Selecting
the optimum number of windings then depends on the available voltage, cur-
rent, and the type of trajectory chosen to reach the required valve dynamics.
For our trajectories, a higher initial jerk leads to lower energy consumption.
It is expected then, that a lower winding number improves efficiency at the
expense of higher currents.
For the FFVA system with a QB02302 motor, Figure 3.3 shows a plot of
maximum current, maximum torque and energy consumption vs. winding
ratio: This plot shows that fewer windings require moderately less energy
and a slightly smaller maximum torque. However, these advantages are
offset by a large increase in required current. Note that there is a minimum
value for maximum current and the winding ratio chosen for the FFVA
system is close to that minimum value.
Before presenting an implementation of a system with the stock QB02302
motor, the following chapter derives how to design a custom motor for this
application.

28
3.4. Design Example

Figure 3.3: Winding variation for QB02302

29
Chapter 4

Motor Optimization

The last chapter showed that motor parameters play an important role in
the energy consumption and the acceleration capability of the FFVA sys-
tem. Motor selection criteria were presented that can be used to choose
the best off-the-shelf motor. However, in the FFVA system, the motor is
used for a transient application, while the majority of motors on the market
are designed for continuous application. Thus, off-the-shelf motors usually
do not provide an optimal design for this application. This chapter then
attempts to provide an analytical method to design an optimum motor.
Several critical constraints should be considered when such a BLDC motor
is designed. First it will be demonstrated that the stator saturation deter-
mines the thickness of the permanent magnets used on the rotor. Then a
constraint on magnet demagnetization will provide a relationship between
the acceleration of the actuator and the motor length. Thereafter, the need
for a linear relationship between rotor movement and valve movement will be
used to define radius of the rotor. Finally, it will be shown, that a relation-
ship between rotor and the stator radius governs the energy consumption of
the actuator.

4.1 Stator Saturation Constraint

The iron core of the stator consists of teeth and slots. The slots are filled
with windings that create a stator field. The teeth are used to guide the
stator field due to the current in the stator windings and the rotor field due
to permanent magnets on the rotor. In order to ensure efficient use of the
stator material, the stator core should not be saturated. This section shows
that stator saturation can be used to determine an optimal thickness of the
permanent magnets on the rotor.
A variety of magnet materials are available to provide the required
rotating magnetic field in a brushless dc motor. The most popular are

30
4.1. Stator Saturation Constraint

neodymium-iron-boron (NdFeB) magnets, which are expensive but have the


highest B − H characteristic. The simplest form of brushless dc motor rotor
construction uses a cylindrical shaft and a ferromagnetic rotating structure
with a surface on which magnets are affixed. The magnets are radially
magnetized and the magnet outer surfaces are ground or preformed to be
concentric with the stator inside diameter. The magnets are held in place by
a structural adhesive to prevent radial movement during operation. A mag-
netic field is generated by the magnets and it passes through the magnetic
circuit formed by different parts of the motor.
The flux density in the airgap Bgo due to the magnet alone and the
acceptable linear current load Ks are important factors in motor design.
In this section, we will show how to choose the best value of Bgo . This
is also called selecting the operating point of the magnet. If a magnet with
radial thickness lm is selected, the flux density in the airgap can be expressed
by the magnet residual flux density Br :
lm
Bgo = Br (4.1)
lge

where lge is the effective length of the airgap including the distance from
the stator teeth to the rotor iron core. Usually the effect of slotting is also
included. Approximately, lge = lg + lm .
Ideally speaking, reducing the airgap lg could obtain the same Bgo by
using less permanent magnet material. But in practice, the minimum value
of the airgap is set by mechanical considerations. The usual range of lg is
above 0.5mm.
Choosing the operating point of the magnet to be at the maximum energy
product of the magnet material will result in minimum magnet volume and
magnet cost [35]. However, the resultant average air gap flux density will be
low (about one half of the magnet Br ), and therefore the armature winding
will take a larger portion of the stator volume to provide the larger total
current in order to generate the same torque.
Using a thicker magnet results in a more expensive magnet, but increases
the airgap flux density, reduces the armature current loading, and results
in a better balance of stator lamination iron and stator copper. But the
flux density in the iron core Bc should also be considered before deciding
Bgo . The maximum flux density in the iron core is usually limited in order
to avoid saturation. For the QB02302 motor, the material of the stator is

31
4.2. Demagnetization Constraint

”TRAFOPERM N3”, whose saturation point is about Bs = 2.03T . In order


to leave some margin, it is set conservatively to 1.8T , thus:
wt + ws
Bc = Bgo < 1.8T (4.2)
wt
where ws is the width of each slot and wt the width of each tooth. For most
machines (including our motors), the slotting of the stator is usually made
wt ≈ ws . Thus Bgo ≈ 0.9T is a reasonable choice.
The next section shows how the demagnetization constraint affects the
acceleration capabilities of the motor. This will lead to a discussion on how
to choose the linear current load Ks .

4.2 Demagnetization Constraint

The performance of a BLDC motor used in transient application is mainly


determined by the maximum acceleration it can provide, thus a high linear
current loading is desirable. But the current density in the stator must be
constrained not to demagnetize the permanent magnet. From a demagne-
tization curve, a critical point (BD , HD ) can be found, which should not
be exceeded if demagnetization is to be avoided. With currently available
NdFeB material, a value of about BD = −0.2T can be achieved below a
temperature of 120o C.
In a BLDC motor, the flux density is distributed as in Figure 4.2. The
first plot shows the geometry of a magnet on the rotor. wp is the width of
each pole, wm is the span width of the magnet and lm is the thickness of the
magnet. The second plot shows BI , which is the sinusoidally distributed flux
density created by the stator current. The last plot adds the flux density due
to the magnet alone, Bgo , and BI in order to obtain the total flux density
along the airgap. Clearly, the flux density of any point on this curve should
be kept above BD . If β is the magnet span in mechanical degrees, or simply
defined as β = wm /wp , then:

BI,peak sinβ ≤ Bgo − BD (4.3)

Since the peak value of BI can be expressed as [36]:



2 2rµ0 Ks
BI,peak = (4.4)
plge

32
4.2. Demagnetization Constraint

Figure 4.1: Demagnetization curve [36]

where Ks is the rms linear current density along the stator periphery. No-
tation p represents the number of poles. So the demagnetization constraint
is given by [36]:
plge (Bgo − BD )
Ks ≤ √ (4.5)
2 2rµ0 sinβ

For a PM motor, the torque can be expressed as [36]:

T = 2πr2 lBg1 Ks sinγ (4.6)

where γ is the angular displacement between the fields produced by the


magnet and the stator current. The maximum value is found at sinγ = 1.
Bg1 is the rms value of the fundamental space component of the air gap flux
density due to the magnet. It evaluates to

2 2Bgo sinβ
Bg1 = (4.7)
π

Since we know the maximum linear current density Ks (see Equation

33
4.2. Demagnetization Constraint

Figure 4.2: Flux density due to current and magnet [37]

4.5), then the maximum torque of a 3-phase motor is given by:


2rlplm Br (Bgo − BD )
Tmax = (4.8)
µ0

In practice the stator length l should be an effective length, thus a length


factor kw (since there are windings stretching out of the stator frame at both
ends) should be added. Then the equation above turns into [38]:
2rlplm Br (Bgo − BD )
Tmax = kw (4.9)
µ0

This equation does not include saturation of the iron core. Especially for
small motors, this can lead to substantial errors. Substituting Equation 4.9

34
4.3. Linearity Constraint

and 3.12 into 3.14 provides an expression for maximum valve acceleration
in terms of geometric dimensions and material properties (the expression of
Jm can be found in next section) of the motor [38]:
√ √
Tmax ra 2plm Br (Bgo − BD )kw l
amax = = √ (4.10)
2Jm µ0 πρm r

The equation above shows that the maximum acceleration is propor-


tional to the number of pole pairs p:
2πr
p= (4.11)
3q(ws + wt )
where q is the slot number per pole per phase. For small motors, q = 0.5
is often used, because it provides a good torque to volume ratio. These
motors are called fractional-slot winding motors [39]. In practice, the mini-
mum tooth and slot widths are limited because of manufacturing constraints.
Thus, p is function of rotor radius r. Substituting Equation 4.11 into the
equation above gives:
√ √
2lm Br (Bg − BD )kw l 2πr √
amax = √ = Kα l (4.12)
µ0 πρm r 3q(ws + wt )

From the equation


√ above, it can be seen that the maximum acceleration
is proportional to l. Thus, for high acceleration requirements, the rotor
length should be chosen as large as possible, or if there is a constraint on
motor length, then the maximum acceleration available can be estimated.
It needs to be pointed out that Equation 4.12 is based on a demag-
netization constraint of the permanent magnets, rather than a saturation
constraint of the iron core. Thus, this equation does not provide good quan-
titative information of the available acceleration, especially for small motors.
It only provides the qualitative guideline that the acceleration is length de-
pendent.The previous sections showed how to select the magnet thickness
and the rotor length. In the upcoming section, criteria for rotor radius are
provided.

4.3 Linearity Constraint

The mechanical optimizations shown in chapter 3 were based on the assump-


tion that there is a linear relationship between valve and rotor motion. This

35
4.4. Energy Constraint

assumption requires that the rotor rotation is small. In practice, one would
usually constrain the rotor rotation to a maximum value of θmax = 30o , thus
the length of the excenter arm is also constrained:
r
Jm s
ra = > (4.13)
m θmax

This defines the lower limit for the rotor’s inertia:


s2 m
Jm > 2
(4.14)
θmax

In practice, the density of the rotor core and the permanent magnets
attached to it, are very similar. Thus, the rotor can be assumed to be a
solid cylinder with density, ρ, motor length, l, and rotor radius, r. The
rotor’s inertia Jm can be expressed as:
1 1
Jm = πr2 lρr2 = πlρr4 (4.15)
2 2

Substituting Equation 4.13 into Equation 4.15 we find:


1 s2 m2
Jm = πlρr4 > 2 (4.16)
2 θmax

So,
2s2 m2
r4 l > 2
(4.17)
θmax πρ

Since the length of the rotor has been determined by the acceleration
requirement of the application, Equation 4.17 provides a constraint for the
minimum rotor radius. At this point the rotor dimensions are fully defined
by the valve acceleration requirements, the constraints on stator saturation,
and linearity between motor and valve motion. The following section shows
how the stator diameter determines the energy consumption of the actuation
device.

4.4 Energy Constraint

In the FFVA system, the energy loss should be less than 2.5J/cycle. When
the motor is designed, the index function JKm2R should be taken into account
T
in order to minimize the energy loss.

36
4.4. Energy Constraint

From above, we know the expression of Jm :


1
Jm = πlρr4 (4.18)
2
and the torque constant KT can be derived as [38]

T 1
KT = = · (Z · 2r · l · Bgo )kw . (4.19)
I 3

The motor’s terminal resistance R can be expressed by a function of


conductors, motor length, and copper area per slot, Aca :

Z 2l
R = f( ) (4.20)
Aca

Then the energy cost term of motor parameter can be rewritten as


1 4 Z2l
Jm R 2 πlρr f ( Aca ) r2
= ∝ , (4.21)
KT2 1
9 · (Z · 2r · l · Bgo · kw )2 Aca

which shows that the cost term is proportional to r2 and A1ca , but is inde-
pendent of the stator length l or number of conductors Z. If we want to
reduce the energy loss, the radius of the rotor should be decreased or the
copper area in each slot increased. The minimum radius is constrained by
the linearity requirement from the previous section. The copper area is con-
strained by the slot area and the ability to fill the slot area with copper. The
remainder of this section is then concerned with the relationship between
stator geometry and copper area in each slot.
The copper area in each slot is related to the wire diameter, dw , and the
number of wires in each slot, nw :
dw 2
Aca = nw π( ) (4.22)
2

The stator slot cannot be completely filled with conductors, since there is
always space left around the wires. The ratio of slot area, Aslot , to conductor
area, Aca , is defined as the filling factor, kf ill :

Aca
kf ill = (4.23)
Aslot

37
4.5. Motor Design Example

The slot area is related to the number of slots, Q, the stator inner radius,
rsi , the slot height, hs , and the tooth width, wt :
π
Aslot ≈ [(rsi + hs )2 − rsi
2
] − wt hs (4.24)
Q

Then the outer radius of the stator is found as:

rso = rsi + (hs + hrs ) (4.25)

where hrs is the required thickness of the stator back. In Figure 2.4, it can
be seen that the stator back must be larger than half the tooth thickness:
1
hrs ≥ wt (4.26)
2

4.5 Motor Design Example

The DC motor (QB02302) chosen for the FFVA system has the parameters
shown in Table 7.1. This is a very good motor, but it was not specifically
designed for our application. This section then shows how to optimize the
motor for the FFVA system.

1. Motor length l: The acceleration capability of the FFVA system is pri-


marily determined by the length of the motor (see Equation 4.12). For
the stock motor, the maximum valve acceleration can be determined
as: amax = Tmax ra 2
2Jm = 6300m/s . The ideal triangular trajectory shown
in Section 3.3 for a lift of 8mm and an engine speed of 6000rpm is ap-
proximately 2500m/s2 . Clearly, the stock motor can provide higher
accelerations than necessary. By reducing the motor length to one
half of its original length, the maximum possible acceleration reduces
to 4400m/s2 . This is still more than the ideal triangular trajectory
prescribes and it leaves room for the suboptimal performance to intro-
duce more inductances.

2. Rotor Radius r: In the last step, the length of the motor was decreased,
thus the rotor inertia Jm becomes smaller. The linearity constraint
needs to be verified. Equation 4.17 can be rewritten as:

2s2 m2
r4 > 2
θmax πρl

38
4.5. Motor Design Example

For lift s = 8mm, valve’s mass m ≈ 40g, motor length l = 28mm the
density of steel ρ = 7800kg/m3 , and θmax = 30o , the limit of r is given
by:
r > 12.8mm
The rotor radius is kept at 14mm, still fulfills the linearity constraint.

3. Slot Area Aslot : The two steps above are used to determine the rotor
size. Energy consumption determines the stator size. It was shown
in Section 3.4 that the energy consumption of the stock system is
expected to be at least 7% larger than required. Thus, the goal of this
step is to reduce the cost term JKm2R of the optimized motor to half of
T
the value of the original QB02302 motor. The energy cost term has not
been affected by the change in rotor length and the parameters left for
us to manipulate are mainly the slot height hs and wire diameter dw .
The number of conductors will be kept. Thus, since the rotor length
was cut in half, the inductance of the motor should also be halved.
This should allow for more energy efficient trajectories.

(a) First, the slot area and related parameters for the original mo-
tor QB02302 are determined. The measured parameters are slot
width ws = 5.1mm, tooth width wt = 5.0mm, number of wires
per slot nw = 72, stator inner radius rsi = 14.5mm, number
of slots Q = 9, airgap lg = 0.5mm, thickness of stator back
hrs = 4.65mm. wire diameter dw = 0.723mm [40] (Gauge 21:
constant maximum current is about 9A, transient maximum cur-
rent is usually 6 ∼ 7 times of that, which is around 50A; the
resistance is 42.0Ohm/km) and slot height hs = 8.5mm.
According to Equation 4.22 the copper area of one slot is given
by:
dw
Aca = nw π( )2 = 29.6mm2
2
Thus, Equation 4.24 provides the expression for the slot area:
π
Aslot = [(rsi + hs )2 − rsi
2
] − wt hs = 68.8mm2
Q
The fill factor can be calculated by using Equation 4.23:
Aca
kf ill = = 43%
Aslot

39
4.5. Motor Design Example

Then outer radius of the stator is found as:

rso = rsi + (hs + hrs ) = 27.65mm

Additionally, the terminal resistance of the motor can be esti-


mated by:
2
R= × 42.0Ohm/km × Zl ≈ 0.254Ohm (4.27)
3
The first constant term 32 was applied because of the relation be-
tween the total terminal resistance (or line resistance) and phase
resistance in the motor. It can be seen that the estimation is
fairly close to the measured value given in Table 7.1, which is
0.24Ohm.
(b) In the next step, the wire diameter and the slot height are ad-
justed in order to provide for lower energy consumption. The slot
height and the wire diameter are set to new values: hs = 14mm,
dw = 1.15mm (Gauge 17: constant maximum current is about
19A, transient maximum current is usually 6 ∼ 7 times of that,
which is around 120A; the resistance is 16.6Ohm/km). Now fol-
lowing the same calculation procedure.

Aca = 74.8mm2

Aslot = 140mm2
kf ill = 53%
rso = 35mm
Thus, the copper cross section area and slot area have both in-
creased. The filling factor becomes a little higher, but it is still in
a reasonable range. The outer radius has increased by 7.35mm.
Most importantly, the cost term JKm2R has decreased. From Equa-
T
tion 4.21, we know the copper loss of the system for an FFVA
system with constant rotor radius is inversely proportional to the
copper area:
Jm R 1
E∝ ∝ (4.28)
KT2 Aca
29.6mm2
Which means theoretically, the copper loss now is 74.8mm2
=
39.6% of the value that QB02302 consumes.

40
4.5. Motor Design Example

Also, the terminal resistance of the motor can be estimated by


2
R = × 16.6Ohm/km × Zl ≈ 0.050Ohm. (4.29)
3
However, the end effect of winding becomes more dominant. Since the
motor length is reduced, the portion of noneffective winding which is
stretching out of the stator increases, which means the actual value of
terminal resistance will be bigger than the calculation.

In this chapter, major constraints of designing an optimized motor for


the FFVA system were studied and a design procedure was presented. The
procedure is systemically shown in the Figure 4.3.

Figure 4.3: Diagram of optimization procedure

The derivation was verified by numerical calculation of the QB02302 pa-


rameters. Finally, the parameters of the QB02302 motor were optimized
using approximate analytical expressions in order to provide better perfor-
mance and a shorter motor. In the next chapter, the predicted performance
of the optimized motor will be verified using finite element calculations.

41
Chapter 5

Finite Element Analysis of


DC Motor

In the FFVA system, a brushless DC motor drives the mechanical part


and its copper loss is responsible for most of the energy losses of the ac-
tuation system. In a conventional internal combustion engine, the energy
loss for each intake valve is about 2.5J/cycle. Right now, the experimental
FFVA platform using a stock QB02302 motor has an energy loss of about
2.7J/cycle. In order to make the system compare favorably with the conven-
tional ones, the energy loss must be reduced. According to the theoretical
derivation, the energy cost term JKm2R should be minimized by optimizing
T
motor parameters. In the last chapter, an optimized motor was given by
derivation and calculation. In this chapter, both the stock QB02302 mo-
tor and the optimized motor are simulated using the finite element analysis
software Maxwell SV. Then the parameters of the two models are compared.
This chapter first introduces the different topologies of the motors. Then
the theoretical background of the FEM calculation is briefly explained. Fi-
nally the results of simulating the two motors are presented.

5.1 Brushless DC Motor Topology

Because of the motor’s symmetry along the axis of the shaft, a 2D model
is sufficient to analyze the important properties, while the complexity of
the calculations are reduced. The transverse sections are given in Figure 5.1
and Figure 5.2. The structures of both motors are also symmetrical by every
120o , only one third of the layout is simulated, which is sufficient and concise
for the Finite Element Analysis. Comparing the two models, it can be seen
that the rotor structure is kept the same and the stator structure is changed
by varying the slot height and outer diameter of the stator.

42
5.2. Maxwell Equations

Figure 5.1: QB02302 Figure 5.2: Optimized motor

5.2 Maxwell Equations

In electromagnetism, Maxwell’s equations are a set of four partial differential


equations that describe the properties of the electric and magnetic fields.
Here, one of them is mainly used, which is called Ampere’s circuital law:


− −
→ → − −
→ → dD − →
I Z Z
HdL = J dS + dL (5.1)
L S S dt

− →

Here, H is the vector of the magnetizing field, J is the vector of the free


current density and D is the vector of the electric displacement field. The
constitutive equations are as follows:

→ →

D = εE

− →

B = µH

− →

J = σE


− →

where E is the vector of the electric field, B is the vector of the magnetic
field, ε is the electric permeability, µ is the magnetic permeability and σ is
the electric conductivity.

43
5.3. Governing Equations in 2D Electromagnetic Field

5.3 Governing Equations in 2D Electromagnetic


Field

From the equations given above, 2D electromagnetic field governing equation


[41] can be derived:

− 1−→ →

∇ × ( B ) = Jt (5.2)
µ

Or,
∂ 1 ∂Ax ∂ 1 ∂Ay
( )+ ( ) + Jz = 0 (5.3)
∂x µ ∂x ∂y µ ∂y
where Jt is the total surface current while Jz is the component of Jt in the


z-direction, and the flux density vector B is defined in terms of the potential


vector A using the auxiliary equation:
− −
→ → −→
∇×A =B (5.4)

The boundary conditions are


 −
→ −

 Γ : 1 ∂ A1 + 1 ∂ A2 = J

µ1 −
→ µ2 →

t s
n n (5.5)
 → −
− →
Γu : A = A0

The natural boundary condition is represented by Γt while the essential


boundary condition is Γu . The equivalent surface current density Js equals
the summation of change rate of the potential vector along the normal vector
to the interface of the two fields, while the subscript refers to different parts
on each side of the boundary. − →
n is the normal unit vector. The potential
vector on the border equals to the prescribed value. As long as we have a
strong form of the governing equations for the 2D electromagnetic field, the
weak form can be derived by using the minimal potential energy principle or
the Galerkin method. Then, the weak form can be applied to each meshed
element to obtain the stiffness matrix and the damping matrix.
So far, the general equations for a 2D electromagnetic field analysis have
been defined. But for different parts of the BLDC motor, the forms of the
equations are different.

44
5.3. Governing Equations in 2D Electromagnetic Field

In the laminated stator, the iron core and air gap are

 Ω1 : ∂ ( 1 ∂Ax ) + ∂ ( 1 ∂Ay ) = 0

∂x µ ∂x ∂y µ ∂y (5.6)
Γ : −
 →
A =0
u

Here, Ω1 represents the field in the stator. Jz becomes 0 since there is


no free current going through the core and the eddy current can be omitted
since the core is formed by stacked silicon steel. Beyond the inner and outer
boundary of the core, the potential vector is 0.
In the rotor iron core,
 ∂ 1 ∂Ax ∂ 1 ∂Ay ∂A
 Ω2 : ∂x ( µ ∂x ) + ∂y ( µ ∂y ) = σ ∂t


−−−→ −−→ (5.7)
 Γt1 : 1 ∂ A− core 1 ∂ Aair

+ = Js1

µ →
n µ →
−n
core air

The term σ ∂A
∂t exists in the solid rotor iron core where eddy currents
cannot be ignored.
There are two models which are commonly used to represent perma-
nent magnets: the magnetization vector method and the equivalent current
sheet method. The magnetizing vector method is used in this model for the
permanent magnets:
 ∂ 1 ∂Ax ∂ 1 ∂Ay 1
 Ω3 : ∂x ( µ ∂x ) + ∂y ( µ ∂y ) = ∇ × ( µ Br )


−−−→ −−→ (5.8)
 Γt2 : 1 ∂ A− PM 1 ∂ Aair

+ = Js2

µ →
n µ →
−n
PM air

Here, the free current density is not 0 anymore, because there is induced
current in the permanent magnet when the magnetic field is changing. The
linear current density between air and the permanent magnet gives the nat-
ural boundary condition.

45
5.4. Finite Element Simulation for BLDC Motor

In the windings, the following conditions hold:



∂ 1 ∂Ax ∂ 1 ∂Ay NI
Ω4 : ( )+ ( )=−


∂x µ ∂x ∂y µ ∂y S





 ∂ 1 ∂Ax ∂ 1 ∂Ay NI
Ω5 : ( )+ ( )= (5.9)

 ∂x µ ∂x ∂y µ ∂y S
Z Z
di N L ∂A ∂A



 U = RI + L +
 (− dΩn + dΩp )
 dt S n ∂t p ∂t

Since there is free current going through the windings in the positive
or the negative direction, the current density can be expressed as the total
current over area. The third equation is the coupling of the field and the
external drive circuit. The boundary conditions were already given in the
previous three parts.
These are governing equations for different parts of motors. They are not
directly solved by the FEA software, because they are all partial differential
equations. They have to be transferred to a weak form, which then can be
solved numerically by the software.

5.4 Finite Element Simulation for BLDC Motor

The previous sections described the motor geometry, the governing equa-
tions, and the boundary conditions of the motor. These models are imple-
mented in Maxwell SV using Galerkin’s method for 3-node elements. This
section describes the meshing process, the Galerkin’s formulations, and typ-
ical plots of flux density and current density that are obtained with Maxwell
SV.
Mesh generation for the FEM should be simple and robust, and the rotor
mesh should be allowed to rotate easily. In this approach, the FEM mesh of
the cross section of the BLDC motor is divided into three parts: the stator,
the rotor and the magnet, with each including a part of the air gap. When
the rotor is rotated according to the time step, the shape of the mesh for
both the stator and rotor can be kept constant and only the coordinates of
the rotor mesh and the periodic boundary condition on the interface need
to change. Therefore, in this approach the stator mesh and the rotor mesh
are required only need to be generated once. This can greatly reduce the
computing time required to generate the FEM mesh at each time step. A

46
5.4. Finite Element Simulation for BLDC Motor

Figure 5.3: Mesh result

typical mesh for the stock QB02302 motor is shown in Figure 5.3. It is
automatically generated by the software, but a finer mesh along the air gap
was specified manually, because this is the most important part of the flux
distribution. Additionally, the software checks for convergence and suggests
finer meshes, if the convergence criteria are not met.
Galerkin’s method is usually employed for the finite element formula-
tion. This method uses particular weighted residuals for both the weighting
functions and the shape functions.
According to the Galerkin’s method, for a 3-node triangular element, the
magnetic vector potential can be expressed as [42]
3
X
A= N i Ai (5.10)
i=1

Here A is the function of the vector potential, Ni is the element shape

47
5.4. Finite Element Simulation for BLDC Motor

function and the Ai is the approximation of the vector potential at the nodes
of the element.
The Galerkins formulation of the laminated stator iron core and air gap
ZZ 3 3
∂Ni ∂ 1 X ∂Ni ∂ 1 X
( N j Aj + Nj Aj )dxdy = 0 (5.11)
∂x ∂x µ ∂y ∂y µ
j=1 j=1

In matrix form, this turns into


1
[ [G]{A}] = 0 (5.12)
µ

For the rotor iron core, the formulation is


ZZ 3 3
∂Ni ∂ 1 X ∂Ni ∂ 1 X ∂Ai
( Nj Aj + Nj Aj + σNi )dxdy = 0 (5.13)
∂x ∂x µ ∂y ∂y µ ∂t
j=1 j=1

In matrix form, this can be written as


1 ∂A
[ [G]{A} + σ[T ] ]=0 (5.14)
µ ∂t

For the permanent magnet, the formulation can be expressed as


ZZ ZZ
1 ∂A ∂Ni ∂A ∂Ni µ0 ∂Ni ∂Ni
( + )dxdy = (Mx − My )dxdy (5.15)
µ ∂x ∂x ∂y ∂y µ ∂y ∂x

In matrix form
[G]A = Brx [ci ] − Bry [bi ] (5.16)

For the winding, the equation is


ZZ 3 3
∂Ni ∂ 1 X ∂Ni ∂ 1 X NI
( N j Aj + N j Aj + N i )dxdy = 0 (5.17)
∂x ∂x µ ∂y ∂y µ S
j=1 j=1

Again, the equation can be rewritten in matrix form:


1 NI
[ [G]{A} + {Q} ]=0 (5.18)
µ S

48
5.5. Evaluation of Optimized Motor

For the circuit equation


Z Z
L ∂A ∂A di
V = [ Ni dΩn − Ni dΩp ] + Ri + L (5.19)
S n ∂t p ∂t dt

In matrix form:
L ∂A ∂A di
V = [({Q}{ })n − ({Q}{ })p ] + Ri + L (5.20)
S ∂t ∂t dt
where, ZZ  ∆e
6 if i = j
Tij = Ni Nj dxdy = ∆e
12 6 j
if i =
ZZ
∆e
Q= Ni dxdy =
3
ZZ
∂Ni ∂Nj ∂Ni ∂Nj bi bj + ci cj
G= ( + )dxdy =
∂x ∂x ∂y ∂y 4∆e
and ∆e is the triangular area of the element.
The field equations above and the circuit equation have to be solved
simultaneously. Rotor movement has to be coupled. The coefficient matrix
is symmetric.
The computed steady-state flux distributions when maximum current
100A is applied in the stator winding is shown in Figure 5.4 (QB02302),
which indicates that the iron core is not saturated even maximum current
is applied. Forces and torques are calculated by integrating the Maxwell’s
stress tensors along a closed path in the air gap.

5.5 Evaluation of Optimized Motor

The simulations in the last section need to be compared to the analytical


derivations from Chapter 4. The properties of the stock and the optimized
motors are shown as in Table 5.1. The parameters of the optimized mo-
tor are given by both the FEM method and the analytical method. In the
simulation, eddy currents and displacement currents are neglected. The
permeability is linear. The iron core material is homogeneous. These as-
sumptions are widely used in modeling DC and AC machines, permanent
magnet devices, transformers and other electrical equipments [43]. In the
table, it can be seen that the optimized motor has a smaller energy cost

49
5.5. Evaluation of Optimized Motor

Parameter QB02302 Optimized Optimized


(FEM) (Theory)
Outer Radius of Stator [mm] 27.65 35 35
Length of stator core [mm] 56 28 28
Wire Diameter [mm] 0.7229 1.15 1.15
Torque Constant KT [Nm/A] 0.076 0.042 0.036
Terminal Resistance Rc [Ohms] 0.24 0.067 0.050
Motor Inertia Jm [kg.m2 ] 2.3E-5 1.5E-5∗ 1.5E-5
Energy Cost Term JK m Rc
2 9.56E-6 5.60E-6 5.79E-6
T

Table 5.1: Comparison of motor parameters

term, which means that it consumes less energy for completing the same
transition. Also, there is good agreement between the FEM simulation and
the analytical expressions. The parameters of size are inherently the same
since the models are identical. The difference of energy cost term between
FEM and theory is mainly caused by two parameters: motor resistance and
torque constant. The motor resistance given by the software includes the
extra windings stretching out of the stator, which is not estimated in our
theocratical calculation. The torque constant in the analytical calculation
is an estimated value, with some assumptions, such as the flux density in
the iron core is uniform and the magnet thickness is even. The FEM model
is likely closer to the true motor parameters.
The value of the energy cost term in the optimized motor is almost half
of that of the stock QB02302. Thus, the energy loss for the same operation is
cut down to half in theory. After optimizing the geometry of the stator, the
quantity of material used in the laminated core is increased. Even though
the core loss has actually increased, it is relatively small comparing to the
copper loss. It can be seen that the overall performance of the optimized
motor is far better than the stock motor QB02302.
This chapter used a finite element simulation in order to validate the
theoretical predictions of the optimized motor. This was achieved by sim-
ulating the original QB02302 motor and the new optimized motor. The
QB02302 motor compared well with the specification sheet and the opti-
mized motor compares well with the theoretical predictions from Chapter
4. The parameters obtained with the FEM model for the optimized motor
will be used in Chapter 7 in order to simulate the overall actuation system.

50
5.5. Evaluation of Optimized Motor

Figure 5.4: Flux density distribution of the optimized motor

51
Chapter 6

Control Strategy

In this chapter, the technology for controlling the FFVA system is discussed.
The main function of the control system is to move the valve from the closed
to the opened position (and vice-versa) avoiding noise, which is caused by
nonzero seating velocity. Similar to typical modern servo control systems, a
cascaded control architecture is employed (see Figure 6.1). A conventional 3
phase inverter provides PI current control and commutation using measured
values of position and current as well as reference current. The reference
current is provided by a PD+Feedforward position controller that are im-
plemented in a dSpace rapid prototyping system. The position controller
receives its reference position from an energy minimized trajectory genera-
tor that is also implemented in dSpace. The algorithms for the trajectory
generator have been provided in Chapter 3. Positions measurements are
provided by an encoder. In order to minimize noise when deriving velocity
from the position signal, a state observer is employed. A detailed description
of the different parts of the controller follows in this chapter.

Figure 6.1: Diagram of control units

52
6.1. 3 Phase Inverter

6.1 3 Phase Inverter

The 3 phase inverter provides commutation and current control. Commuta-


tion is the process of switching current to generate a rotating magnetic field.
In brushless systems, commutation is accomplished electronically using a ro-
tor position feedback device. The rotor position sensors may be magnetic,
optical, or any type of device that provides sufficiently accurate and reliable
information on the motor position. In general, the best position for commu-
tation is that point at which the back EMF waveform is ”centered” between
commutation points.
In addition to commutation, the 3 phase inverter also provides PI current
control. Thus, to the outside the 3 phase inverter combined with the BLDC
motor looks like a single phase DC motor plus inverter, because the 3 phase
inverter takes on the function of the mechanical commutation device found
in a conventional single phase DC motor. It is thus possible to model the
3 phase inverter and brushless DC motor as an equivalent single phase DC
motor. A block diagram of this representation together with the PD position
controller is shown in Figure 6.2. The position controller is built around the
current controller. Usually in cascaded control loops, the inner current loop
has a bandwidth ωi , that is considerably higher than the bandwidth ωp , of
the outer position control loop. This allows independent tuning of the two
control loops. Then, the current controller is shown in Figure 6.3:
Ignoring the back EMF due to motor speed, the transfer function of the
current controller shown in Figure 6.3 can be written as
Pi Ki
I(s) Pi s + K i Ls+ L
= = (6.1)
Iref (s) Ls2 + s(R + Pi ) + Ki s2 + s( R+P
L )+
i Ki
L

The proportional gain Pi and the integral gain Ki of the PI controller


can be derived from desired bandwidth of the current controller, ωi (usually,
ζ is close to 1):
Ki Pd + R
ωi2 = ; 2ζωi =
L L
Then,
Pi = 2ζωi L − R; Ki = ωi2 L

The value of the ωi is usually selected to be approximately five times


higher than the bandwidth of the position controller.

53
6.2. Position (or PD) Controller

6.2 Position (or PD) Controller

Position control can be implemented in a variety of different ways. Typically,


Either a Lead lag controller or a PD controller are used for this task. In our
case we have chosen a PD controller because it provides a simple strategy to
set bandwidth and damping for the position loop. It also does quite well in
rejecting disturbances and it fairly robust towards model uncertainties. Once
the gains of the PI controller are determined, the current can be regulated
fast and accurately. For the position controller, the current controller can
be idealized as the unit gain block shown in Figure 6.4.
The transfer function (Bm can usually be neglected which reduces the
complexity of the transfer function) for position control can be written as
Kd Pd
θ(s) K d s + Pd J s+ J
= = (6.2)
θref (s) Js2 + Kd s + Kd s2 + KJd s + PJd

The proportional gain Pd and the derivative gain Kd of PD controller


can be derived from the mechanical resonance frequency ωp (usually, ζ is
close to 1):
Pd Kd
ωp2 = ; 2ζωp =
J J

Then,

Pd = ωp2 J; Kd = 2ζωp J

The value of ωp needs to be sufficiently high in order to allow tracking of


the reference trajectory. For the FFVA system the trajectories at 6000rpm
contain a fundamental frequency component at approximately 100Hz with
very few harmonics. Thus it is sufficient to set the bandwidth of the position
controller to 350Hz or ωp = 700 × πrad/sec.

6.3 States Observer

The PD position controller requires measured values of position and velocity.


In commercial servo applications, resolvers or encoders are commonly used
for measuring position. Velocity is either found by digitally differentiating

54
6.3. States Observer

the position signal or by employing a tachometer. For the FFVA application


in a real engine, encoders would not be suitable, because they are too ex-
pensive and cannot withstand the extended temperature range found in the
cylinder head. Inductive pickups or hall sensors are much likelier solutions.
However, in the test bed used in this thesis, an encoder is used, because it
is readily available and provides excellent position resolution as well as high
bandwidth.
The principal operation of the optical encoder is explained here. The en-
coder consists of light sources, a slotted disk, photo detectors, a bearing-unit
and signal processing circuits. After the signal light passes through the slits
on the the rotating disk, it is captured by photo detectors, and transduced
to an electronic signal. There are two types of disks, an incremental and an
absolute type disk. However, creating high resolution absolute encoders is
more difficult and most encoders are of the relative type.
The encoder chosen for the FFVA test bed has a resolution of 20, 000 lines
per revolution, which provides a highly accurate measurement of position. In
Simulink, this encoder can be modeled by a ”Quantizer” with a quantization
interval of 2π/80000.
The PD controller requires a position and a velocity measurement. To
avoid additional hardware it is desirable to derive the velocity measurement
from the position measurement. Even though the resolution of the direct
position measurement is relatively high, the differentiation operator ampli-
fies high frequency noise and reduces the resolution by the inverse of the
sampling time:

resolution(θ)
resolution(ω) =
SamplingT ime
resolution(ω)
resolution(α) =
SamplingT ime

Since the typical sampling time is 0.025ms, the resolution for velocity
and acceleration using a simple backward difference operator deteriorates
drastically. The quantization noise on the measured velocity is small for
the encoder type chosen in the experimental setup. However, real sensors
in actual engine applications would usually provide a much lower resolution
that would then lead to significant quantization errors in the derived velocity
measurements. A low pass filter (LPF) can be introduced to reduce this noise

55
6.3. States Observer

and quantization errors, although it introduces a delay. The cutoff frequency


of the LPFs should be high compared to the bandwidth of the position loop
since this delay may deteriorate the performance of the control system.
In order to get better velocity estimation, a Kalman filter can be used.
The specific Kalman filter designed for FFVA application is shown here.
The transfer function of a DC motor is generally described by
KT
θ(s) (Jm s+bm )(Ls+R) 1
= KT
(6.3)
e(s) 1 + KB (Jm s+bm )(Ls+R) s

The numerical values for the parameters of the DC motor is given in


Table 7.1.
When we design a discrete Kalman Filter for the system, discretization
[44] should be done. The state vector is defined as:
 
θ
q = α 
ω

If we set the sampling time at 0.25e−4 ms, and use the parameters given
above, the discrete state space can be calculated:
   
1 0.001529 0 0
A = 0 0.9794 0  B = 1.0e−4  0 
0 121.9 0.9999 0.001529
 
C= 1 0 0 D=0

The predictor corrector type procedure of the Kalman filter can then be
written as
q(k) = q̄(k) + G(k)[y(k) − Cq̄(k))] (6.4)
q̄(k + 1) = Aq(k) + Bu(k) (6.5)
T T −1
G(k) = M(k)C [CM(k)C + Rv ] (6.6)
P(k) = M(k) − G(k)CM(k) (6.7)
M(k + 1) = AP(k)AT + Rw (6.8)
where q̄(k) is the predicted state estimate at the sampling instant k, and
q(k) is the actual state estimate. G(k) is the Kalman gain. M(k) is the

56
6.3. States Observer

covariance of the prediction errors. Matrices Rv and Rw are the covariance


matrices of the observation noise, and disturbance signals respectively.
The covariance of the input disturbance in a real engine cylinder are not
estimated in our research. A large numerical value is used in order to ensure
that sufficient robustness towards disturbances is achieved. The actual value
of the covariance of the input disturbance was determined by observing the
control stability of the actuator when exposed to different types of friction
forces.
Generally, the estimation of Rv is based on the assumption that it is true
random noise with a Gaussian form and a standard deviation comparable
to the size of the least significant bit of the sensor. Then the matrix has the
following form:
Rv = cov(θ) = [E(θ2 )] (6.9)

For sensor quantization, the error probability for any one sample is uni-
form within the region −0.5 to 0.5, and zero elsewhere. The covariance of
the noise for one sample can be calculated as:
Z ∞Z ∞
Rv = xpθ (x) × xpθ (x)dxdx
−∞ −∞
Z 0.5 Z 0.5
= xpθ (x) × xpθ (x)dxdx (6.10)
−0.5 −0.5
2(0.5)3
= = 0.0833
3
where, pθ (x) is the function of probability.
The Kalman Filter is simulated in Simulink together with other system
components. Simulation results comparing the performance of different fil-
tering techniques are given in chapter 7. They show the advantage of the
Kalman Filter over direct differentiation and low pass filtering.

57
6.3. States Observer

58
Figure 6.2: Cascaded control structure
6.3. States Observer

Figure 6.3: Current controller

59
6.3. States Observer

Figure 6.4: Position controller

60
Chapter 7

System Simulation and


Experimental Validation

In the previous chapters, two setups were introduced. The first setup uses
a stock QB02302 motor, and the second setup uses an optimized custom
motor that provides lower power consumption.
The first part of this chapter describes the components of an experimen-
tal test bed used in this study. The second part of this chapter compares
the transition performance of the two motors using a Simulink simulation
for typical engine operating scenarios. The third part of this chapter looks
at the robustness of the actuation system towards parameter variations of
the mechanical design and variations of the motor parameters. The fourth
part of this section focusses on the ability of the control system to reject
external disturbances due to combustion pressures or friction. Finally, the
fifth part of this chapter describes the experimental results obtained with
the stock QB02302 setup.

7.1 Experimental Setup

The experimental test bed drives a single exhaust valve of a Honda cylinder
head (see Figure 7.1). A block diagram of the FFVA system is shown in
Figure 6.1. It contains a dSpace1103 controller board that creates trajecto-
ries and performs position control. For position measurements, a 20000 line
encoder from Quantum devices is used. The control board performs its tasks
at 40kHz in order to match the PWM frequency of the subsequent 3 phase
inverter from Maccon GmbH. The inverter performs current control and
commutation using high speed PWM controlled power MOSFETs that can
regulate up to 100A per phase. The inverter internally uses LEM modules
to feed the current controllers. In addition, three external LEM HTP100-

61
7.2. Simulation of Transition Performance

Figure 7.1: System hardware

P current probes are employed to measure and display currents with the
dSpace system. The inverter drives a QB02302 motor that is mounted to a
small cylinder head from Honda. The light weight excenter arm is fabricated
in aluminum and has a length of 23mm in order to match the 37g valve to
the inertia of the motor. The excenter arm and the valve are joined using a
connecting rod made from 2mm piano wire. Table 7.1 lists the parameters
(QB02302 with customized winding) of the FFVA system.

7.2 Simulation of Transition Performance

Before building an experimental setup, the overall system is simulated using


Simulink. All mechanical, electrical and magnetic components of the system
are included in this simulation.

62
7.2. Simulation of Transition Performance

QB02302 with customized winding


Torque Constant KT 0.056 Volts
Back EMF constant KB 0.056 V/rad/s
Peak Current Ip 100 Ampere
Terminal Resistance R 0.13 Ohms
Terminal Inductance L 0.15 mH
Wire Diameter dw 0.7229 mm
Motor Inertia Jm 2.3E-5 kg.m2
Airgap Flux Density Bg ∼ 0.9 Tesla
Magnet Thickness lm 3.5 mm
Effective Airgap lge 4.05 mm
Stator Inner Radius rsi 14.5 mm
Stator Outer Radius rso 27.5 mm
Motor Length l 56 mm
Number of Phases 3
Number of Poles 6
Number of Slots 9
Number of Conductors per Slot 36

Table 7.1: Motor Parameters

Two separate simulations are performed. First, the stock motor QB02302
system is simulated and its performance is estimated. Second, the optimized
motor is simulated in order to compare its possible performance improve-
ment with the predictions from the design procedure.
The transition performance of the closed loop controlled system can be
quantified by using the following three indices [45]:
Transition Time: The time it takes for the valve to move from 5% of the
maximum lift to 95% of the maximum lift.
Valve Seating Velocity: The velocity of the valve when it contacts the
valve seats.
Energy Loss: The copper loss per valve per transition.
The initial simulation of the actuation system with the stock QB02302
motor is based on the parameters listed in Table 7.1.
A number of different operating conditions are simulated to demonstrate
the flexibility and performance of the FFVA system. Figure 7.2 shows the

63
7.3. Simulation of the Robustness Towards Parameter Variations

simulated valve lift curves and Figure 7.3 depicts the corresponding energy
consumptions.
The dark blue curve shows a typical 8mm valve lift curve for 6000rpm,
which corresponds to t2 = 2.4ms. As predicted in Section 3.4, this tra-
jectory requires 1.35J/transition which is slightly more than the desired
1.25J/transiton. This problem can be alleviated by reducing the lift. A
trajectory with half the lift and the same transition time is shown in green
and it requires approximately 0.35J which is less than a third of the energy
consumed for twice the lift.
In practice, the required lift reduction would be much smaller in order to
achieve the desired energy consumption. The red curve shows an 8mm lift
curve for 3000rpm. Since the additional valve open time was achieved by
inserting extra time with the valve completely open, rather than reducing
the valve accelerations, the energy consumption is still approximately 1.35J
in every transition. In practice, one would likely reduce the acceleration in
order to find a compromise between engine performance and valve energy
loss. The light blue curve represents a 4mm valve lift curve at 8000rpm.
Compared to the case of 6000rpm with the same lift, this mode requires
almost three times the energy. This highlights the nonlinear relationship
between energy consumption and transition time.
Even though the setup with the stock QB02302 motor performs quite
well in the simulations, the energy consumption at high rpm is slightly
larger than desired. The optimized custom motor was designed to address
this problem in Chapter 4. Figure 7.4 shows the displacement curves and
energy consumption for simulating the optimized motor at 6000rpm en-
gine speed. The energy consumption for this setup has been reduced from
1.35J/transtion to 0.66J/transition. This corresponds well with the pre-
dictions shown in Table 5.1.

7.3 Simulation of the Robustness Towards


Parameter Variations

The system parameters such as valve mass m, excenter arm length ra , motor
torque constant KT and winding resistance R can shift from the nominal
values due to manufacture accuracy, abrasion, temperature or other factors.
In the simulation, the values of the listed four parameters are all increased

64
7.3. Simulation of the Robustness Towards Parameter Variations

by 30% from the nominal value. Each parameter is varied individually in


one simulation.
The results of these simulations are shown in Figure 7.5 and they demon-
strate that the control system is able to compensate for a wide range of
parameter variations, since all the motion curves including position, veloc-
ity and acceleration are still accurately following the reference trajectories.
However, energy losses are different. The sensitivity of the energy con-
sumption towards parameter variations can be estimated by taking partial
derivatives of Equation 3.10 with respect to the parameters being varied.
For the parameters studied here this leads to:
∂E
δE ∂m δm δm
≈ = (7.1)
E E m
∂E
δE ∂R δR δR
≈ = (7.2)
E E R
∂E
δE ∂Kt δm δKt
≈ = −2 (7.3)
E E Kt
∂E
δE ∂ra δra
≈ =0 (7.4)
E E
Note that since the optimum excenter arm length was chosen to provide
minimum energy consumption, the derivative of the energy with respect to
arm length and hence the sensitivity in Equation 7.4 need to be zero. It is

parameter stationary estimated sensitivity simulated


δE
point E sensitivity
Valve mass m 0.037g 30% 38%
Motor resistance R 0.13Ohm 30% 31%
Torque constant Kt 0.056Nm -60% -42%
Excenter arm length ra 0.023m 0 10%

Table 7.2: Sensitivity

expected that actual sensitivity values will differ from the nominal values,
because large changes are made to the parameters and Equations 7.1 to 7.4
provide only a linearized approximation. Table 7.2 shows a comparison of
the estimated and the simulated sensitivity values. It demonstrates that the
torque constant, the resistance, and the valve mass have a large influence on

65
7.4. Simulation of Disturbance Rejection

the energy consumption, whereas the excenter arm length is not very sen-
sitive. This lets us conclude, that manufacturing tolerances of the linkage
system do not need to be very high in order to guarantee low energy con-
sumption. On the other hand all the other parameters need to be designed
carefully.

7.4 Simulation of Disturbance Rejection

The FFVA system is designed for intake valves where only moderate com-
bustion pressure is expected on the valves. The shape of the pressure curve
during the opening or the closing transition is not well defined. Other distur-
bances could include some degree of viscous damping or coulomb damping in
the motor or in the valve guides. In this investigation we attempt to combine
all of these causes using disturbance torques on the motor that are modeled
with a viscous friction coefficient of 0.015N ms/rad and a coulomb coeffi-
cient of approximately 1.15N m. Each of the two parameters corresponds to
an equivalent disturbance force of 50N on the valve.
It should be pointed out, that the original controller was not designed
to compensate for disturbances in the resting positions at either end of
the valve motion. In order to achieve low steady state errors under these
circumstances, the original PD position controller needs to be extended to
a PID controller. The I term is used to reduce steady state error and the
PD terms are defined in the same fashion as in the original design.
Figure 7.6 shows simulation results that indicate that the system is able
to cope with significant disturbance forces. The additional energy required
by the actuation system is very close to the energy dissipated in friction.
We conclude, that the efficiency of the actuation system in the presence of
disturbances is still close to optimal.

7.5 Simulation of Low Resolution Sensor with a


Kalman Filter

The experimental system used in this thesis uses a very high resolution op-
tical encoder. In a real engine, this is not a suitable sensor, because it is
too expensive and cannot withstand the harsh environmental conditions. A
hall sensor is a more likely candidate, and it would usually have a resolu-

66
7.6. Experimental Results

tion somewhere between 7 and 10 bits. At the high sampling rates used
for this system, the low resolution would lead to considerable quantization
error, that would consequently introduce large noise in the position signal.
For eliminating measurement noise we will now compare a low pass filter
and a Kalman filter. The low pass filter is a simple backward difference
implementation of a continuous first order lag with a break frequency that
is ten times higher than the position loop bandwidth. The Kalman filter
implementation and the selection of the covariances for the disturbances
and the measurement follow the guidelines shown in Chapter 6. Figure 7.7
shows the systems response using an 7 bit sensor with 1 bit measurement
noise and different feedback strategies. When the sensor signal and a di-
rect backwards differentiator are used, the system is unstable. When the
sensor and the backwards differentiator are combined with low pass filters
with bandwidths at 10 times the bandwidth of the position controller, the
oscillations are reduced but not eliminated. Only the Kalman filter, is able
to provide a stable response for the low resolution sensor.
The reason for the difference in control performance can be found by
comparing position and velocity errors for the different feedback signals.
Figure 7.8 then shows this comparison. As expected, the position error of
the Kalman filter is slightly smaller than the actual sensor signal. However,
the low pass filter demonstrates considerable phase lag which manifests itself
in a position error proportional to velocity. The velocity error of the direct
backward difference differentiator is the same size as the full scale velocity
measurement. This error is reduced considerably by both the low pass filter
and the Kalman filter. However, the latter shows less than half the noise
levels of the low pass filter.
In summary, for a low resolution sensor system, the Kalman filter is the
only approach investigated here that guarantees sufficient robustness.

7.6 Experimental Results

In this section, the experimental system performance of the stock QB02302


motor is compared with the simulations and with literature data from EMCV
and EMVD systems. In these experiments, the valve opens 8mm in approx-
imately 3.4ms, which corresponds to a typical valve motion at a 6000rpm
engine speed.
Figure 7.9 shows position, velocity and acceleration of the valve. Track-

67
7.6. Experimental Results

Lif t = 8mm EMCV1 EMCV2 EMVD FFVA


[46] [47] [14]
Transition 3.5 3.5 3.4 3.4
Time [ms]
Seating 0.3 0.1 0.15∼0.27 0.1∼ 0.25
Velocity [m/s]
Energy Loss 3.0∼5.2 - 1.4 1.35(QB02302)
[J/transition] 0.66(Optimized)
Voltage 100 42 42 42
Source [V ]
Controllable Duration Duration Duration Fully Variable
Variables & Phase & Phase & Phase & Programmable

Table 7.3: Performance comparison

ing performance is quite good. However, due to the flexibility in the linkage
between valve and motor, mechanical vibrations are induced. This cause
was determined by comparing the standard operation to a setup with an
equivalent inertia mounted rigidly to the motor. With the rigid inertia,
no noticeable vibration was observed. This leads to the conclusion that
the rather rudimentary linkage implementation shown here needs to be re-
designed to provide less flexibility in future setups. The velocity plot shows
that the mechanical vibrations due to the flexible link are primarily excited
during the high jerk periods at the beginning and at the end of the motion.
Consequently, the seating velocity at the end of the valve travel is rather
high. It should be noted that the transition time chosen here would be used
during high rpm of the combustion engine. However, the low valve seating
velocity of 0.05m/s is only required during engine idle, when the transition
times are much longer and the valve lift can be reduced significantly. Under
such circumstances, the FFVA system will provide valve seating velocities
well below the desired value.
Phase current, total current, and energy loss are plotted in Figure 7.10.
The motor draws approximately 100A during acceleration and 60A during
deceleration. The unsymmetrical behavior and the large current oscilla-
tions during accelerations are again attributed to the flexible linkage sys-
tem, which has unsymmetrical damping elements for opening and closing.
However, the energy consumption of 1.35J/transition in the experiment
corresponds well to the analytical predictions in Chapter 4. Table 7.3 shows

68
7.6. Experimental Results

a comparison of the performance characteristics of the FFVA system with


the EMVD and the EMCV system. For a similar transition time and lift,
the FFVA system performs as good or better than the other two systems.
However, the FFVA system provides not only variable valve timing but also
variable lift. This additional flexibility leads to significant operating advan-
tages for the combustion engine.

69
7.6. Experimental Results

Figure 7.2: Simulation of displacement trajectories

70
7.6. Experimental Results

Figure 7.3: Simulation of energy losses of different trajectories

71
7.6. Experimental Results

Figure 7.4: Transition simulation of the optimized motor


72
7.6. Experimental Results

Figure 7.5: Simulation of parameter variations


73
7.6. Experimental Results

Figure 7.6: Simulations of disturbance rejection


74
7.6. Experimental Results

Figure 7.7: Simulation of transition response with a 7 bit sensor and various
filtering strategies

75
7.6. Experimental Results

Figure 7.8: Simulations of measurement error using different filtering strate-


gies

76
7.6. Experimental Results

Figure 7.9: Experimental trajectory results using QB02302

77
7.6. Experimental Results

Figure 7.10: Experimental energy loss using QB02302

78
Chapter 8

Conclusion

8.1 Summary

In this thesis, a Fully Flexible Valve Actuation system for the intake valves
of automotive internal combustion engines is demonstrated.
After reviewing the need for electronic valve actuation and current imple-
mentations of such systems, a direct drive system is proposed that provides
six main advantages:

• Fully flexible valve control

• Simple mechanical layout

• An effective control system

• Good energy efficiency: 1.35J/transition using QB02302 (simulation


results show that the energy loss is only 0.66J/transition using an
optimized motor)

• Fast transition: sufficient for the operation at 6000rpm engine speed

• Relatively low seating velocity, which is regulated to about 0.2m/s at


6000rpm and has the potential to be improved. Also, at engine idle
the system can provide the desired seating velocity of 0.05m/s.

For a successful implementation, all aspects of the actuation system need


to be optimized. This includes the mechanical portion, the electromag-

79
8.2. Suggestions for Future Works

netic portion, and the control algorithms of the actuation system. A design
methodology for an actuation system based on stock motors is proposed
first. It provides guidelines how to select a good motor, how to design the
mechanical layout, and how to design an energy efficient control strategy. In
the course of this work, analytical expressions were developed to predict the
transition performance and the energy efficiency of the actuation system. In
addition, a lumped parameter model was developed to support the analyt-
ical performance predictions. The lumped parameter model was also used
to study the influence of varying system parameters, the ability to reject
disturbances, and effects of sensors with low resolution. Finally a test bed
was constructed that was used to validate the transition performance. These
investigations show that the FFVA system based on stock motors perform
well compared to other systems found in the literature.
To further improve the performance of the FFVA system, an analytical
design methodology for modifying the stock motors is then proposed. An
FEM model is used to validate the motor optimization methodology and
the lumped parameter model is applied again in order to predict the system
performance. This study shows that the modified motor should provide
superior energy efficiency without compromising transition speed.

8.2 Suggestions for Future Works

Even though the current experimental setup performs well compared to


other systems found in literature, there are still a number of areas of im-
provement:
As shown in Figure 7.9, the seating velocity is still relatively high. This
is likely due to the elasticity of the connecting rod that causes a bounce
when the valve comes to the end of a transition. A stiffer joint needs to be
designed, or the control system needs to be refined in order to reduce the
seating velocity.
A second area of work is the experimental implementation of the op-
timized custom motor. The theocratical derivation and simulation of the
optimized motor have been presented in this thesis, and this motor promises
to improve the performance of the actuation system significantly. However,
the scope of the thesis did not allow for an experimental implementation.
Finally, the current system is powered by a voltage source of 42V . For

80
8.2. Suggestions for Future Works

better compatibility with current vehicles, the supply voltage needs to be


decreased to 12V . Further optimization has to be performed on the current
system structure to meet this requirement.

81
Bibliography

[1] M. Forissier, “Reducing greenhouse gases,” Forum Annuel de lIn-


vestissement Responsable en Europe, sponsored by French SIF, Tech.
Rep., June 2007.

[2] N. Milovanovic, R. Chen, and J. Turner, “Influence of variable valve


timings on the gas exchange process in a controlled auto-ignition en-
gine,” in Proceedings of the Institution of Mechanical Engineers Part
D-Journal of Automobile Engineering, vol. 218, 2004, pp. 567–583.

[3] R. R. Henry and B. Lesquesne, “A novel, fully-flexible, electro-


mechanical engine valve actuation system,” SAE Tech. Paper Series,
1997.

[4] S. Butzmann, J. Melbert, and A. Koch, “Sensorless control of electro-


magnetic actuators for variable valve train,” SAE Tech. Paper Series,
2000.

[5] E. Sher and T. Bar-Kohany, “Optimization of variable valve timing


for maximizing performance of an unthrottled si engine - a theoretical
study,” Energy, 2002.

[6] R. R. Chladny, C. R. Koch, and A. F. Lynch, “Modeling automotive


gas-exchange solenoid valve actuators,” IEEE Transactions On Mag-
netics, vol. 41, no. 3, March 2005.

[7] J. G. Calvert, J. B. Heywood, R. F. Sawyer, and J. H. Seinfeld, “Achiev-


ing acceptable air quality: Some reflections on controlling vehicle emis-
sions,” Science, vol. 261, no. 5117, pp. 37–45, July 1993.

[8] M. Pischinger, W. Salber, F. V. D. Staay, H. Baumgarten, and H. Kem-


per, “Low fuel consumption and low emissions electromechanical valve
train in vehicle operation,” International Journal of Automotive Tech-
nology, vol. 1, no. 1, pp. 17–25, September 2000.

82
Chapter 8. Bibliography

[9] M. Theobald and B. L. and.R. Henry, “Control of engine load via elec-
tromagnetic valve actuators,” SAE Tech., vol. 103, pp. 1323–1334, 1994.

[10] V. A. W. Hillier and F. W. Pittuck, Fundamentals of motor vehicle


technology. Hutchinson, 1966.

[11] W. Hoffmann and A. G. Stefanopoulou, “Valve position tracking for


soft landing of electromechanical camless valvetrain,” Advances in Au-
tomotive Control 2001, pp. 295–300, 2001.

[12] W. S. Chang, T. A. Parlikar, M. D. Seeman, D. J. Perreault, J. G.


Kassakian, and T. A. Keim, “A new electromagnetic valve actuator,”
Ieee Power Electronics in Transporation, pp. 109–118, 2000.

[13] K. Peterson, A. Stefanopoulou, T. Megli, and M. Haghgooie, “Output


observer based feedback for soft landing of electromechanical camless
valvetrain actuator,” in Proceedings of the 2002 American Control Con-
ference, Vols 1-6, 2002, pp. 1413–1418.

[14] T. A. Parlikar, W. S. Chang, Y. H. Qiu, M. D. Seeman, D. J. Perreault,


J. G. Kassakian, and T. A. Keim, “Design and experimental implemen-
tation of an electromagnetic engine valve drive,” IEEE/ASME Trans-
actions On Mechatronics, vol. 10, no. 5, October 2005.

[15] C. F. Taylor, The internal-combustion engine in theory and practice.


M.I.T. Press, 1985.

[16] A. Nix, “The future of 42v vehicle systems,” Vehicle Systems Integra-
tion in the Wired World, pp. 167–176, 2001.

[17] H. T. and H. M., “Development of the variable valve timing and lift
(vtec) engine for the honda nsx,” SAE transactions, vol. 100, no. 3, pp.
1–7, 1991.

[18] Moriya, Y. Saitoh, T. Kawatake, and M. K. Yoshioka, “Vvt-i system,”


TOYOTA TECHNICAL REVIEW, vol. 47, no. 1, pp. 48–53, 1997.

[19] J. S. Brader, “Development of a piezoelectric controlled hydraulic ac-


tuator for a camless engine,” Master’s thesis, Boston University, 1995.

[20] Z. Sun and D. Cleary, “Dynamics and control of an electro-hydraulic


fully flexible valve actuation system,” in Proceedings of the American
Control Conference, Denver, Colorado, June 2003, pp. 3119–3124.

83
Chapter 8. Bibliography

[21] C. Tai, T.-C. Tsao, and M. B. Levin, “Adaptive nonlinear feedforward


control of an electrohydraulic camless valvetrain,” in Proceedings of the
American Control Conference, June 2000.
[22] P. K. Wong, L. M. Tam, and K. Li, “Modeling and simulation of a dual-
mode electrohydraulic fully variable valve train for four-stroke engines,”
International Journal of Automotive Technology, vol. 9, pp. 509–521,
2008.
[23] J. M. Donaldson, “Dynamic simulation of an electrohydraulic open cen-
ter gas-change valve actutor system for camless internal combustion
engines,” Master’s thesis, Milwaukee School of Engineering, 2003.
[24] S. K. Chung, C. R. Koch, and A. F. Lynch, “Flatness-based feedback
control of an automotive solenoid valve,” IEEE Transactions On Con-
trol Systems Technology, vol. 15, no. 2, March 2007.
[25] D. O. Wright and J. A. Nitkiewicz, “Engine valve actuation control
system,” United States Patent, no. 5632, 1999.
[26] R. Seethaler, J. Meyer, A. Knaut, and K.-H. Gaubatz, “Pivoting actua-
tor for controlling the stroke of a gas exchange valve in the cylinder head
of an internal combustion engine,” United States Patent, no. 11/128185,
2006.
[27] M. GmbH, “Quantum series brushless servo motors,” http:// www.
maccon.de/ FTPROOT/ Quantum.pdf , 2009.
[28] Brushless Motors and Drive Systems, Kollmogen Inland Motor.
[29] R. W. J. Armstrong, “Load to motor inertia mismatch:unveiling the
truth,” in DRIVES AND CONTROLS CONFERENCE, Telford Eng-
land, 1998.
[30] A. Sasane, “Calculus of variations and optimal control,” September
2004.
[31] A. M. Trzynadlowski, “Energy optimization of a certain class of incre-
mental motion dc drives,” IEEE Transactions On Industrial Electron-
ics, vol. 35, no. 1, February 1988.
[32] Y. A. Kaan Erkorkmaz, “High speed cnc system design. part iii: high
speed tracking and contouring control of feed drives,” International
Journal of Machine Tools and Manufacture, vol. 41, no. 2001, p.
1637C1658, October 2000.

84
Chapter 8. Bibliography

[33] Pittman, “64mm series brushless dc motor,” http:// www.ametektip.


com/ index.php? option=com rokdownloads&view=file&id=1454 ,
2009.

[34] MCG, “Motor i2383092nc-xxxx,” http:// www.ametektip.com/ index.


php? option=com content&view=article&id=192 , 2009.

[35] R. H. Engelmann and W. H. Middendorf., Handbook of Electric Motors.


Dekker, 1995.

[36] G. Slemon, “On the design of high-performance surface-mounted pm


motors,” IEEE Transactions on Industry Applications, vol. 30, no. 1,
pp. 134–140, Jan/Feb 1994.

[37] T. Sebastion and G. R. Slemon, “Transient modeling and performance


of variable speed permanent magnet motors,” IEEE Transactions on
Industry Applications, vol. 25, no. 1, Jan/Feb 1989.

[38] Y. Chin, W. Arshad, T. Backstrom, and C. Sadarangani, “Design of a


compact bldc motor for transient applications,” in European Coference
on Power Electronics, Austria, 2001.

[39] N. Bianchi and S. Bolognani, “Fractional-slot pm motors for electric


power steering systems,” International Journal of Vehicle Autonomous
Systems, vol. 2, no. 3-4, pp. 189–200, 2004.

[40] P. Technology, “Wire gauge and current limits,” http:// www.


powerstream.com/ Wire Size.htm, 2009.

[41] Y. Zhang, K. T. Chau, J. Z. Jiang, D. Zhang, and C. Liu, “A finite


elementcanalytical method for electromagnetic field analysis of electric
machines with free rotation,” IEEE Transactions On Magnetics, vol. 42,
no. 10, 2006.

[42] J. P. A. Bastos and N. Sadowski, Electromagnetic Modeling By Finite


Element Methods. Marcel Dekker, Inc., 2003.

[43] M. Chari, G. Bedrosian, and J. D’Angelo, “Finite element applications


in electrical engineering,” IEEE TRANSACTIONS ON MAGNETICS,
vol. 29, no. 2, 1993.

[44] G. R. Yu, M. H. Tseng, and Y. K. Lin, “Optimal positioning control of


a dc servo motor using sliding mode,” in Proceedings of the 2004 IEEE
International Conference on Control Applications, 2004.

85
Chapter 8. Bibliography

[45] R. Pierik and J. Burkhard, “Design and development of a mechanical


variable valve actuation system,” in SAE technical Paper Series, 2000.

[46] Y. Wang, T. Megli, M. Haghgooie, K. Peterson, and A. Stefanopoulou,


“Modeling and control of electromechanical valve actuator,” Society of
Automotive Engineers, vol. 01, no. 1106, 2002.

[47] R.R.Chladny and C.R.Koch, “Flatness-based tracking of an electrome-


chanical variable valve timing actuator with disturbance observer feed-
forward compensation,” IEEE transactions on Control System Tech-
nology, vol. 16, no. 652-633, 2008.

86

You might also like