Lattimore 16 PH D
Lattimore 16 PH D
by
Thomas Lattimore
A thesis submitted to
DOCTOR OF PHILOSOPHY
School of Engineering
December 2015
i
University of Birmingham Research Archive
e-theses repository
This research has examined the combustion and emissions of a spray-guided direct-injection
spark-ignition (DISI) engine using exhaust gas recirculation (EGR). The impact of EGR type,
swirl and tumble intake airflows, compression ratio and fuel type were also investigated.
EGR addition resulted in significant fuel consumption improvements and differing particulate
matter (PM) behaviour depending on the knock limited maximum brake torque (KLMBT)
spark advance achieved. When comparing EGR types, cooled EGR achieved the best fuel
consumption and cooled EGR after three-way catalyst (TWC) achieved the best gaseous
Swirl and tumble intake airflows significantly increased fuel consumption. However, these
increases could be minimized with EGR addition. Swirl significantly reduced the
EGR addition did not significantly affect PM for the swirl and tumble intake airflow
conditions.
8.5 bar IMEP. At 7.0 bar IMEP, EGR addition allowed the KLMBT spark timing to be
advanced, as the compression ratio was increased. Fuel consumption was improved by 0.4%
due to the spark advance and reduced pumping losses, and PM improved because the
ii
Acknowledgements
First and above all, I praise God, the most beneficent, the most merciful, for providing me
with this opportunity and for granting me the capability to succeed. This thesis would not
have been possible without the assistance and guidance of several people. I would therefore
I would like to give my warmest and most sincere gratitude to Professor Hongming Xu who
throughout the duration of my PhD program has provided me with invaluable guidance,
experience and support to enable me to successfully complete my PhD. I would like to thank
my second supervisor, Professor Miroslaw L. Wyszynski for his useful comments and
feedback for my research work, and I would like to thank Professor Akbar Ghafourian for his
guidance, help and encouragement in my annual reviews. I would also like to thank the
Engineering and Physical Sciences Research Council for providing the funding for my PhD
scholarship, and I would like to thank the University Of Birmingham for facilitating the
scholarship and my PhD studies, as well as Professor Kyle Jiang for awarding the scholarship
to me.
Great thanks go to my PhD candidate and research assistant colleagues who have provided
me with technical knowledge and a good working environment to complete my PhD studies.
I would like to thank Dr. Chongming Wang, Dr. Jose Herreros, Soheil Rezaei, Dr. Changzhao
Jiang, Tawfik Badawy, Dr. Xaio Ma, Dr. Dhananjay Srivastava, Dr. Isaline Lefort, Dr.
Arumugam Sakunthalai Ramadhas, Dr. Mohammadreza Attar, Dr. He Ma and Maria Macias,
in particular, for their technical support at various times during my studies. I would also like
to thank Yasser Al Qahtani, Rafiu Olalere, Ricky Creegan and Lucas Polglase for providing a
iii
I am grateful to Navin Kalian from Jaguar Land Rover for his technical support during the
PhD and to Shell Global Solutions UK for their supply of fuels to the engine labs. I am also
grateful to the Automotive Laboratory staff who have provided significant technical support,
knowledge and expertise. This is in addition to manufacturing and installing new components
for the engine test cell, and maintaining the Future Engines and Fuels Laboratory at the
University Of Birmingham. They are Carl Hingley, Peter Thornton, Lee Gauntlett and Jack
Garrod.
Finally, I would like to thank my wife, Mariam Kamel for her endless love, support and care
during my PhD studies, and our families, in particular our parents, for their love and
iv
Contents
Abstract ................................................................................................................................ ii
Acknowledgements .............................................................................................................. iii
Contents ................................................................................................................................ v
List of Figures ..................................................................................................................... vii
List of Tables ........................................................................................................................ x
List of Notations .................................................................................................................. xi
List of Publications ............................................................................................................. xv
Chapter 1 Introduction .......................................................................................................... 1
1.1 Overview ..................................................................................................................... 1
1.2 Objectives and Approaches.......................................................................................... 3
1.3 Research Outline ......................................................................................................... 4
1.4 Thesis Outline ............................................................................................................. 5
Chapter 2 Literature Review.................................................................................................. 7
2.1 Introduction ................................................................................................................. 7
2.2 History of IC Engines .................................................................................................. 8
2.3 SI Combustion ............................................................................................................. 9
2.4 Spark Ignition Fuels ................................................................................................... 22
2.5 Regulated Engine-Out Emissions ............................................................................... 29
2.6 Summary ................................................................................................................... 37
Chapter 3 Experimental Setup and Techniques .................................................................... 39
3.1 Introduction ............................................................................................................... 39
3.2 Overview of Single-Cylinder Engine Test Cell .......................................................... 40
3.3 Control Systems ......................................................................................................... 57
3.4 Instrumentation .......................................................................................................... 61
3.5 Emissions Measurement ............................................................................................ 66
3.6 Data Acquisition and Processing ................................................................................ 73
3.7 Analysis of the Uncertainties in the Recorded Data .................................................... 91
3.8 Fuel Properties ........................................................................................................... 92
3.9 Experimental Test Procedures .................................................................................... 93
3.10 Summary ................................................................................................................. 94
v
Chapter 4 The Effect of EGR and its Type on Engine Combustion and Emissions ............... 95
4.1 Introduction ............................................................................................................... 95
4.2 Experimental Procedure ............................................................................................. 96
4.3 Results and Discussion .............................................................................................. 98
4.3.1 Hot EGR Effect on Combustion and Emissions in a DISI Engine ......................................... 98
4.3.2 Effect of Different EGR Types on Combustion and Gaseous Emissions in a DISI Engine .. 124
4.4 Data Continuity ....................................................................................................... 147
4.5 Conclusions ............................................................................................................. 148
Chapter 5 The Effect of Intake Airflow and Hot EGR on Engine Combustion and PM
Emissions .......................................................................................................................... 151
5.1 Introduction ............................................................................................................. 151
5.2 Experimental Procedure........................................................................................... 152
5.3 Results and Discussion ............................................................................................ 154
5.4 Data Continuity ....................................................................................................... 175
5.5 Conclusions ............................................................................................................. 175
Chapter 6 The Effect of Compression Ratio, Fuel and EGR on Engine Combustion and
Emissions .......................................................................................................................... 177
6.1 Introduction ............................................................................................................. 177
6.2 Experimental Procedure ........................................................................................... 179
6.3 Results and Discussion ............................................................................................ 181
6.3.1 Investigation of Compression Ratio and Fuel Effect on Combustion and Emissions ........ 181
6.3.2 Investigation of the Compression Ratio and Hot EGR Effect on Combustion and PM
Emissions............................................................................................................................... 204
6.4 Data Continuity ....................................................................................................... 217
6.5 Conclusions ............................................................................................................. 218
Chapter 7 Summary, Conclusions and Suggestions for Future Work ................................. 221
7.1 Summary and Conclusions ....................................................................................... 221
7.2 Suggestions for Future Work ................................................................................... 225
Appendix A1: Analysis of the Uncertainties in the Recorded Data .................................... 228
Appendix B1: Data Continuity (Chapter 4)........................................................................ 230
Appendix B2: Data Continuity (Chapter 5)........................................................................ 232
Appendix B3: Data Continuity (Chapter 6)........................................................................ 234
Appendix C1: Detailed Specific Further Investigation for Each Research Area ................. 236
List of References ............................................................................................................. 240
vi
List of Figures
Figure 3-1 Schematic of single-cylinder engine test facility and instrumentation setup .... 40
Figure 3-2 Schematic of water cooling and oil lubrication circuits ................................... 44
Figure 3-3 Schematic of fuel supply systems ................................................................... 45
Figure 3-4 Relative positions of the direct injector along with the injector spray plume ... 47
Figure 3-5 Cooled EGR schematic .................................................................................. 49
Figure 3-6 Hot EGR after TWC schematic ...................................................................... 50
Figure 3-7 Cooled EGR after TWC schematic ................................................................. 52
Figure 3-8 a) Low tumble baffle plate, b) swirl baffle plate, c) high tumble baffle plate .. 52
Figure 3-9 ETCS LabView program front-panel (Luszcz, 2009)...................................... 57
Figure 3-10 VVT LabView program front-panel (Luszcz, 2009) ....................................... 59
Figure 3-11 Intake and exhaust camshaft profiles .............................................................. 60
Figure 3-12 Engine ETCS and VVT control systems; their arrangement and signal flow ... 61
Figure 3-13 Calibration results of the VAF Meter ............................................................. 65
Figure 3-14 Cambustion DMS 500 analyser operating principle ........................................ 68
Figure 3-15 Schematic of the DMS 500 two-stage dilution system .................................... 69
Figure 3-16 HSDAQ LabView program front-panel (Luszcz, 2009) .................................. 74
Figure 3-17 LSDAQ LabView program front-panel (Luszcz, 2009) .................................. 75
Figure 3-18 Ricardo WAVE simulation model .................................................................. 76
Figure 3-19 Ricardo WAVE model calibration procedure ................................................. 78
Figure 3-20 In-cylinder pressure versus CAD data comparisons for Ricardo WAVE model
calibration: a) case 1, b) case 2, c) case 3, d) case 4 and e) case 5 ........................................ 80
Figure 3-21 Engine data comparison for Ricardo WAVE model calibration: a) IMEP (bar),
b) Pmax (bar), c) indicated efficiency and d) VE ................................................................... 81
Figure 3-22 Engine data percentage difference comparison for Ricardo WAVE model
calibration: a) IMEP, b) Pmax, c) indicated efficiency and d) VE .......................................... 82
Figure 3-23 In-cylinder pressure versus CAD data comparisons for Ricardo WAVE model
verification: a) case 1, b) case 2, c) case 3, d) case 4 and e) case 5 ....................................... 84
Figure 3-24 Engine data comparison for Ricardo WAVE model verification: a) IMEP (bar),
b) Pmax (bar), c) indicated efficiency and d) VE ................................................................... 85
Figure 3-25 Engine data percentage difference comparison for Ricardo WAVE model
verification: a) IMEP, b) Pmax, c) indicated efficiency and d) VE ......................................... 86
Figure 3-26 Modified HSDAQ LabView program displaying the on-line knock
amplitudes…… ................................................................................................................... 90
Figure 4-1 In-cylinder pressures versus CAD at KLMBT spark timings for a) 5.5 bar
IMEP, b) 7.0 bar IMEP and c) 8.5 bar IMEP ..................................................................... 103
Figure 4-2 Calculated average in-cylinder temperatures versus CAD at KLMBT spark
timings for a) 5.5 bar IMEP, b) 7.0 bar IMEP and c) 8.5 bar IMEP .................................... 105
Figure 4-3 MFB versus CAD at KLMBT spark timings for a) 5.5 bar IMEP, b) 7.0 bar
IMEP and c) 8.5 bar IMEP ................................................................................................ 109
Figure 4-4 Gravimetric ISFCnet versus EGR ratio at KLMBT spark timings .................. 112
vii
Figure 4-5 EGT versus EGR ratio at KLMBT spark timings ......................................... 114
Figure 4-6 PN emissions versus particle diameter at KLMBT spark timings at a) 5.5 bar
IMEP, b) 7.0 bar IMEP and c) 8.5 bar IMEP ..................................................................... 119
Figure 4-7 NOx emissions versus EGR ratio at KLMBT spark timings .......................... 121
Figure 4-8 HC emissions versus EGR ratio at KLMBT spark timings ........................... 123
Figure 4-9 In-cylinder pressures versus CAD at KLMBT spark timings for a) hot EGR, b)
cooled EGR, c) hot EGR after TWC and d) cooled EGR after TWC.................................. 127
Figure 4-10 Calculated average in-cylinder temperatures versus CAD at KLMBT spark
timings for a) hot EGR, b) cooled EGR, c) hot EGR after TWC and d) cooled EGR after
TWC…………………..………………………………………………………………...…..130
Figure 4-11 MFB versus CAD at KLMBT spark timings for a) hot EGR, b) cooled EGR,
c) hot EGR after TWC and d) cooled EGR after TWC ...................................................... 133
Figure 4-12 MFB50 versus EGR ratio at KLMBT spark timings ................................. 134
Figure 4-13 MFB10-90 versus EGR ratio at KLMBT spark timings ............................ 135
Figure 4-14 Gravimetric ISFCnet versus EGR ratio at KLMBT spark timings............... 138
Figure 4-15 PMEP versus EGR ratio at KLMBT spark timings ................................... 140
Figure 4-16 EGT versus EGR ratio at KLMBT spark timings ...................................... 141
Figure 4-17 COVIMEP versus EGR ratio at KLMBT spark timings ............................... 142
Figure 4-18 Intake plenum temperature versus EGR ratio at KLMBT spark timings .... 143
Figure 4-19 NOx emissions versus EGR ratio at KLMBT spark timings ...................... 145
Figure 4-20 HC emissions versus EGR ratio at KLMBT spark timings ........................ 147
Figure 5-1 In-cylinder pressure verses CAD at KLMBT spark timings for a) the three
intake airflow conditions at 0% EGR, b) low tumble, c) swirl, d) high tumble ................... 159
Figure 5-2 Calculated average in-cylinder temperature verses CAD at KLMBT spark
timings for a) the three intake airflow conditions at 0% EGR, b) low tumble, c) swirl, d) high
tumble………… ............................................................................................................... 162
Figure 5-3 MFB verses CAD at KLMBT spark timings for a) the three intake airflow
conditions at 0% EGR, b) low tumble, c) swirl, d) high tumble ......................................... 166
Figure 5-4 Gravimetric ISFCnet versus EGR ratio at KLMBT spark timings .................. 168
Figure 5-5 EGT versus EGR ratio at KLMBT spark timings ......................................... 169
Figure 5-6 PN emissions versus particle diameter at KLMBT spark timings for a) the three
intake airflow conditions at 0% EGR, b) low tumble, c) swirl, d) high tumble ................... 173
Figure 6-1 In-cylinder pressure versus CAD for a) Bu20, b) E20 and c) ULG95 at KLMBT
spark timings.. ................................................................................................................... 185
Figure 6-2 Calculated average in-cylinder temperature versus CAD at KLMBT spark
timings for a) Bu20, b) E20 and c) ULG95 ........................................................................ 187
Figure 6-3 MFB versus CAD at KLMBT spark timings for a) Bu20, b) E20 and c)
ULG95……….. ................................................................................................................ 189
Figure 6-4 MFB50 versus compression ratio at KLMBT spark timings ......................... 190
Figure 6-5 MFB10-90 versus compression ratio at KLMBT spark timings .................... 191
Figure 6-6 Indicated efficiency versus compression ratio at KLMBT spark timings....... 193
Figure 6-7 EGT versus compression ratio at KLMBT spark timings .............................. 194
Figure 6-8 PN emissions versus particle diameter at KLMBT spark timings for a) Bu20, b)
E20 and c) ULG95 ............................................................................................................ 197
viii
Figure 6-9 NOx emissions versus compression ratio at KLMBT spark timings .............. 201
Figure 6-10 HC emissions versus compression ratio at KLMBT spark timings ............ 202
Figure 6-11 Overall effect of compression ratio and fuel on gaseous emissions, indicated
efficiency and total PN (integrated across 10-289 nm range) at KLMBT spark timings ..... 203
Figure 6-12 In-cylinder pressure versus CAD at KLMBT spark timings for a) 0% EGR, b)
7% EGR and c) 14% EGR................................................................................................. 207
Figure 6-13 Calculated average in-cylinder temperatures versus CAD at KLMBT spark
timings for a) 0% EGR, b) 7% EGR and c) 14% EGR ....................................................... 209
Figure 6-14 MFB versus CAD at KLMBT spark timings for a) 0% EGR, b) 7% EGR and
c) 14% EGR….. ................................................................................................................ 212
Figure 6-15 Gravimetric ISFCnet versus compression ratio at KLMBT spark timings ... 213
Figure 6-16 PN emissions versus particle diameter at KLMBT spark timings for a) 0%
EGR, b) 7% EGR and c) 14% EGR ................................................................................... 217
ix
List of Tables
Table 2-1 EU (EN 228-2012) and US (ASTM D4814 Rev B-2011) Fuel Specification
(Analiit AA, 2015, Chevron, 2009) ..................................................................................... 23
Table 2-2 Operating conditions for research and motor octane number test methods
(Bradley, 2004, Heywood, 1988)......................................................................................... 24
Table 2-3 European Union emission legislation for DISI petrol cars (Delphi, 2012) .......... 35
Table 2-4 United States Federal emission legislation for Light Duty SI Vehicles <80,000 km
(Delphi, 2012) ..................................................................................................................... 36
Table 3-1 Experimental single-cylinder engine specification ............................................. 41
Table 3-2 Estimated swirl and tumble ratios for the three intake airflow conditions ........... 53
Table 3-3 Summary of other researchers findings for swirl and tumble ratios .................... 54
Table 3-4 Single-cylinder engine camshaft geometry......................................................... 59
Table 3-5 SMPS 3936 system settings ............................................................................... 71
Table 3-6 Horiba MEXA-7100DEGR specification ........................................................... 72
Table 3-7 Test cases for Ricardo WAVE model calibration ............................................... 79
Table 3-8 Test cases for Ricardo WAVE model verification .............................................. 82
Table 3-9 Test Fuel Properties ........................................................................................... 92
Table 3-10 Gasoline GC Analysis ....................................................................................... 93
Table 4-1 Experiment test matrix (EGR addition & engine load) ....................................... 98
Table 4-2 Experiment test matrix (EGR addition & EGR type) .......................................... 98
Table 4-3 KLMBT spark timings (˚bTDC) (* = not knock limited) (EGR addition & engine
load)……… ...................................................................................................................... 100
Table 4-4 KLMBT spark timings (˚bTDC) (EGR addition & EGR type) ......................... 125
Table 4-5 Results summary (EGR addition & engine load) (highlighted=improvement,
underlined=worsening)...................................................................................................... 149
Table 4-6 Results summary (EGR addition & EGR type) (highlighted=improvement,
underlined=worsening)...................................................................................................... 150
Table 5-1 Experiment test matrix (intake airflow & EGR addition) ................................. 154
Table 5-2 KLMBT spark timings (˚bTDC) (intake airflow & EGR addition) ................... 156
Table 5-3 Results summary (intake airflow & EGR addition) (highlighted=improvement,
underlined=worsening)...................................................................................................... 176
Table 6-1 Experiment test matrix (compression ratio & fuel) ........................................... 181
Table 6-2 Experiment test matrix (compression ratio & EGR addition)............................ 181
Table 6-3 KLMBT spark timings (˚bTDC) (compression ratio & fuel) ............................ 183
Table 6-4 KLMBT spark timings (˚bTDC) (compression ratio & EGR addition) ............. 205
Table 6-5 Results summary (compression ratio & fuel) (highlighted=improvement,
underlined=worsening)...................................................................................................... 219
Table 6-6 Results summary (compression ratio & EGR addition)
(highlighted=improvement, underlined=worsening) .......................................................... 220
x
List of Notations
CC Cubic Capacity
CE Cooled EGR
CI Confidence Interval
CID Combustion Initiation Duration (Crank angle from spark discharge to MFB5)
CO Carbon Monoxide
Comb. Combustion
DC Direct Current
DI Direct-Injection
EC European Commission
xi
EPSRC Engineering and Physical Sciences Research Council
EU European Union
GC Gravimetric Content
HC Unburned Hydrocarbon
HE Hot EGR
IC Internal Combustion
IN Inputs
MF Methylfuran
xii
MFB5 Crank Angle at 5% Mass Fraction Burned
NA Naturally Aspirated
NI National Instruments
NO Nitrous Oxide
η Thermal Efficiency
OUT Outputs
PAI Polyamid-Imide
PM Particulate Matter
PN Particulate Number
Rc Compression Ratio
xiii
ROHR Rate Of Heat Release
SI Spark-Ignition
Sig. Significant
US United States
Vc Clearance Volume
Vd Swept Volume
VE Volumetric Efficiency
λ Air-Fuel Ratio
xiv
List of Publications
Thomas Lattimore, Chongming Wang, Hongming Xu, Miroslaw L. Wyszynski, Shijin Shuai,
Emissions in a DISI Engine, Thomas Lattimore, José M. Herreros, Hongming Xu, Shijin
behaviour and emissions, Chongming Wang, Hongming Xu and Thomas Lattimore. 2013.
Injection Spark Ignition (DISI) Engine, Chongming Wang, Hongming Xu, Jose Martin
Herreros, Thomas Lattimore and Shijin Shuai. 2014. Energy & Fuels, 28, 2003-2012, doi:
10.1021/ef402234z.
Awards:
2013
xv
Chapter 1
Introduction
The aim of this chapter is to provide an overview of the PhD research investigation
conducted by the author. The study of fuel consumption reducing techniques of EGR (also
used to reduce NOx), swirl and tumble intake airflows, geometric compression ratio increase
as well as the study of biofuels is driven by the demand for reduced fuel consumption and
hence reduced CO2 emissions from vehicles. This is in addition to, in the case of biofuels, the
demand for renewable energy supply and increased energy supply security. The overall
research outline, objectives and investigation approach are presented briefly, after which the
1.1 Overview
While many of us dream of a future where electric cars charged by renewable means are
commonplace and the small demand for a traditional engine sound is met by clean hydrogen
combustion, this dream is still a long way away. Crude oil supplies have at least 50 years
remaining (BP, 2015), if not more, especially because of the ongoing discovery and
exploitation of shale oil reserves, which has been a significant factor in the oil price slump we
are experiencing in 2015-16. Therefore, there is not a strong economic incentive for industry
to invest in the technology and infrastructure required for electric and hydrogen powered
vehicles. There are, however, legislative factors that are driving the research and
development in the automotive industry; most significantly, emissions legislation and tax
incentives. The progress though is not towards wide-scale electric vehicle use. It is towards
increased engine and vehicle efficiency through engine downsizing and vehicle hybridization,
amongst other methods, as well as reduced vehicle emissions produced by the efficiency
1
improvements and improved emissions after-treatment. Thus, for the meantime, conventional
spark-ignited and compression-ignited internal combustion (IC) engines are here to stay.
recover waste power from the engine and the vehicle to improve their overall efficiencies.
In spark-ignition (SI) engines, many technologies are being widely utilized to improve fuel
economy and to reduce emissions. These technologies include EGR, improved combustion
using swirl and tumble generated intake airflows, optimised compression ratio, DI,
turbocharging, variable valve lift and timing, cam profile switching and lean-burn stratified
charge combustion. These improve the fuel economy of the engine itself and reduce its
emissions, although PM and some gaseous emissions can potentially be increased. More
efficient catalytic converters and improved particulate filters are being utilized to reduce the
emissions of the vehicle as a whole, by converting the emissions from the engine into less
harmful substances.
Biofuels, particularly bio-ethanol, are also believed to be part of the solution to reduce
vehicle emissions. This is due to their oxygenated nature which reduces combustion
temperatures, providing decreased NOx emissions, along with reduced HC and PM emissions
similar biofuels using renewable methods greatly reduces their net CO2 output; the CO2
produced during combustion is absorbed from the atmosphere when the raw plant material
While EGR, swirl and tumble intake airflows, compression ratio and biofuels have been
studied in detail previously by other researchers, their effect on combustion in DISI engines
and their subsequent emissions has been much less thoroughly investigated, particularly
2
regarding their effect on PM emissions. Therefore, their impact on DISI engine combustion
The main research objective of this investigation was to study the impact of EGR on the
single-cylinder DISI thermal research engine was used for the engine experiments. PM
emissions were measured using a Cambustion DMS 500 and a TSi SMPS 3936. A Horiba
MEXA-7100DEGR gas analyser was used to measure the gaseous emissions of nitrous
The following are the areas that were investigated throughout the author’s PhD study:
The effect of EGR and its type on engine combustion and emissions
The effect of intake airflow and EGR on engine combustion and PM emissions
The effect of compression ratio, fuel and EGR on engine combustion and emissions
Investigation of the effect of different types of EGR (hot EGR, cooled EGR, hot EGR
after TWC, cooled EGR after TWC) on engine combustion and gaseous emissions at
the same engine load using a simulated TWC (this removes the effect of varying
Unlike most other investigation, the PM analysis in this thesis has classified the
particles into nucleation and accumulation modes, which have been studied
3
1.3 Research Outline
The research presented in this thesis was conducted by the author at the Future Engines and
Fuels research laboratory at the University of Birmingham with help and support from Jaguar
Land Rover and Shell Global Solutions UK. The research was mainly conducted to
investigate the effects of EGR on the combustion and emissions of a DISI engine.
The study of engine combustion includes an analysis of the KLMBT spark timing achieved at
different operating conditions with the engine. It includes a detailed and comprehensive
analysis of the in-cylinder pressure trace and mass fraction burned (MFB) profiles. A Ricardo
WAVE engine model has been utilized to calculate the in-cylinder temperature in order to
assist in the engine combustion study. Fuel consumption behaviour has been analysed in
detail to quantify the effect of the combustion at different engine conditions on this
parameter. Pumping mean effective pressure (PMEP), exhaust gas temperature (EGT) and
coefficient of variation of IMEP (COVIMEP) have been analysed too in order to further
quantify the combustion effects. Regulated emissions of PM, NOx and HC were then
examined in order to study the effect of the engine combustion on these critical parameters,
quantified and subsequently shown to be competitive with fuels currently on the market (i.e.
gasoline). Therefore, this thesis also presents research comparing gasoline and oxygenated
4
1.4 Thesis Outline
This thesis is composed of seven chapters. A brief outline of each chapter is presented below.
This chapter reviews the literature that is relevant to this investigation. DI and spark timing
are talked about in detail as well as the phenomenon of engine knock. The current trend of
engine downsizing is introduced along with fuel consumption improvement and CO2
emission reduction, with EGR, swirl and tumble intake airflows and compression ratio
explored in detail. After this, spark ignition fuels including oxygenated alternatives are
discussed. Finally, a discussion into the regulated engine-out emissions and related legislation
is provided.
This chapter gives detailed information of the engine and instrumentation setup, as well as
detailed information of the data acquisition and recording systems. The emission analysers
are introduced after which key calculations for engine parameters used in this thesis are
presented. Finally, the details of the Ricardo WAVE engine model are provided along with
Chapter 4- The Effect of EGR and its Type on Engine Combustion and Emissions
This chapter provides details of the combustion and emissions characteristics of a DISI
engine operated with hot EGR at different engine loads of 5.5, 7.0 and 8.5 bar IMEP, and
with different EGR types of hot EGR, cooled EGR, hot EGR after TWC and cooled EGR
after TWC. In the first part, the hot EGR addition was increased across the different engine
loads, and in the second part the EGR addition was increased at a single engine load of 7.0
5
bar IMEP, and the EGR type is changed. The combustion and emission parameters at the
different conditions are compared to quantify the effect of EGR and its type on them.
Chapter 5- The Effect of Intake Airflow and Hot EGR on Engine Combustion and PM
Emissions
This chapter provides details of the combustion and PM emission characteristics of a DISI
engine operated with swirl and high tumble intake airflows, and with hot EGR addition. The
EGR addition was increased across the two intake airflow conditions along with a baseline
low tumble condition at a single engine load of 7.0 bar IMEP. The combustion and PM
emission parameters at the different conditions are compared to quantify the effect of intake
Chapter 6- The Effect of Compression Ratio, Fuel and Hot EGR on Engine Combustion
and Emissions
This chapter provides details of the combustion and emissions characteristics of a DISI
engine operated with different compression ratios and different fuels, and with hot EGR
addition. In the first part, the two fuel splash blends of Bu20 and E20 were tested along with
a baseline of gasoline at a single engine load of 8.5 bar IMEP, with the compression ratio
increased after each of the fuel blends and the reference fuel were tested. In the second part,
gasoline fuel was tested at a single engine load of 7.0 bar IMEP, with the hot EGR addition
increased along with the compression ratio. The combustion and emission parameters at the
different conditions are compared to quantify the effect of compression ratio, fuel type and
This chapter summarizes the research conducted in this investigation and provides the key
conclusions of this thesis. Recommendations for future work are then given.
6
Chapter 2
Literature Review
The aim of this chapter is to review the literature that is relevant to this investigation. The
2.1 Introduction
Firstly, this chapter provides a brief introduction regarding the history of IC engines and
explains the key fundamental differences between spark and compression ignition engines.
Spark-ignited combustion is then further discussed, after which DI and spark-timing which
result in the combustion and emissions from the engine are talked about in detail. The
phenomenon of knock is then discussed with the factors affecting knocking tendency
explained.
Secondly, the current and ever ongoing trend of improved vehicle fuel economy and reduced
CO2 emissions is introduced followed by detailed discussion of the potential of EGR, swirl
and tumble intake airflows and geometric compression ratio increase to enable these
improvements to be achieved. SI fuels are then explored due to their significant effect on
combustion and emissions, including oxygenated alternatives, due to their increased use in
Finally, regulated engine-out emissions are discussed, including an overview of the current
legislation in both Europe and the US, due to their strong influence on the research and
development of automotive engines. TWCs are then briefly discussed due to their strong
7
2.2 History of IC Engines
The concept of the IC engine can be traced back to 1680 when Dutch Physicist Christian
Huygens designed a gunpowder fuelled IC engine (Stone, 1999, The Institute of Historical
Research, 2004). During the next 200 years further designs were proposed and built to
advance the concept further. The most significant breakthrough came with Nicolaus Otto who
built the first practical four-stroke IC engine in 1876 (Wu, 2007). The Otto cycle formed the
underlying working principle for the engine operation and it became the universally adopted
cycle for all liquid fuelled spark-ignited IC engines to this day. Later in 1893, Rudolph Diesel
achieved compression ignition in an engine operated using the Diesel cycle which is utilized
The fundamental difference between gasoline and diesel engines is not the type of ignition
but the type of combustion that occurs, with spark-ignited engines typically burning with a
pre-mixed flame and compression-ignited engines typically burning with a diffusion flame
(Stone, 1999). In order to obtain a reliable ignition and combustion with a pre-mixed flame,
necessitates the use of a throttle valve to control the amount of air entering the combustion
chamber, which in return increases engine pumping work and thus reduces cycle efficiency
(Stone, 1999). SI engines also suffer from knock, which limits their compression ratio thus
limiting their thermal efficiency (Gupta, 2013). However, in diffusion flames, the fuel-air
mixture is only required to be approximately stoichiometric at the flame front, which allows
the engine to be de-throttled and the fuel injection to be controlled to produce the required
power output (Stone, 1999). Thus, there are no throttling losses, significantly increasing the
economy (Stone, 1999). Also, knock is not a limitation in CI engines, thus, they can be
8
operated with higher compression ratios, significantly improving their thermal efficiency as
engines due to the diffusion flame, necessitating the use of a diesel particulate filter, which
increases pumping losses (Bennett, 2010). The pre-mixed flame produces little PM meaning
the vehicle does not require a particulate trap. Thus, it does not suffer from any additional
pumping losses as a result of the higher exhaust back-pressure (Agarwal et al., 2014).
Today the main goal of engine researchers is to improve engine efficiency in order to reduce
engine CO2 emissions through reduced fuel consumption, while complying with ever tighter
emissions regulations (Zaccardi et al., 2009). This has led to the increased use of DI,
turbocharging and increased compression ratios (Heywood and Welling, 2009, Zaccardi et
al., 2009).
2.3 SI Combustion
During the compression stroke of the SI engine cycle, the piston causes the fuel-air mixture to
become highly compressed before a spark ignites the mixture causing it to combust. The
spark leaves a small nucleus of the flame that propagates into the unburned gas, and it is not
until that nucleus grows into a size which is the same order as that of the turbulence scale that
the flame propagation can be enhanced by in-cylinder turbulence (Stone, 1999). This period
is often referred to as the early burn, delay period or combustion initiation duration (CID) and
it comprises the initial laminar combustion and the transition to fully turbulent combustion,
typically lasting 1-2ms or 9-18 CAD at 1500 rpm (Stone, 1999). The crank angle interval
between the spark discharge and the MFB5 or MFB10 parameters are typically used to define
this period (Stone, 1999). The duration of this period depends on the temperature, pressure
9
Between MFB10 and MFB90 is the main combustion period which is dominated by turbulent
combustion, the second stage of which occurs shortly after the peak pressure and it is affected
by the same parameters as the early burn period along with turbulence (Stone, 1999). The
flame in the final combustion stage makes contact with more of the combustion chamber
walls and consequently it makes less contact with the remaining unburned fuel-air mixture,
2.3.1 Direct-Injection
DI is a technology that was initially developed for diesel engines but it was not until
relatively recently that it was incorporated into gasoline production engines, with the modern
DISI engine being first produced in 1996 by Mitsubishi, although the concept was initially
2013, Mitsubishi Motors, 1999). It has an advantage of improved fuel economy and power
output in comparison to PFI because the fuel can be injected into the combustion chamber at
the best time; it is not dictated by the intake valve opening time as with PFI (Harada et al.,
1997). DI benefits from charge cooling; the fuel cools as it is injected into the combustion
chamber, causing the knocking tendency to reduce, allowing the spark timing to be
Yang and Anderson, 1998). This contributes to fuel economy and power output
improvements at medium-high loads (Alger et al., 2000, Daniel et al., 2012b, Harada et al.,
1997). Both DI and PFI provide an advantage compared to carburettor type fuel injection
because much more precise fuel metering is achieved which improves fuel economy and
reduces engine emissions (Institution of Mechanical Engineers, 2013, Robert Bosch GmbH,
2006). Spray-guided DI first produced by Mercedes in 2006 is the most widely used DI
method due to its improved thermodynamic efficiency and fuel economy compared to other
10
DI systems, such as wall-guided and air-guided DI, in addition to improved HC and PM
homogenous combustion has been used to refer to that produced by PFI and stratified
combustion has been used to refer to that produced by DI. However, within DI itself there are
homogenous and stratified concepts (Harada et al., 1997, Takagi and Skippon, 1998). With an
early injection of fuel directly into the cylinder during the induction stroke, a more
homogenous mixture of approximately λ=1 can be achieved (Harada et al., 1997, Takagi and
Skippon, 1998). When the fuel is injected late into the cylinder during the compression stroke
a more stratified mixture is formed with an overall lambda of ~20 and a combustible mixture
of λ~1 around the spark plug (Harada, 1997, Takagi and Skippon, 1998). With homogenous
mixture formation, the mixture preparation is assisted by the high flow velocities and their
aerodynamic forces in the area of the opening and closing intake valves, helping the fuel to
evaporate quickly, ensuring it is well homogenized (Robert Bosch GmbH, 2006). With
stratified mixture formation, a mixture cloud is created which can then be transported into the
area around the spark plug by the air flows in the combustion chamber and by the upward
stroke of the piston (Robert Bosch GmbH, 2006). The mixture formation benefits from the
higher pressures and temperatures in the combustion chamber during the compression stroke
2.3.2 Spark-Timing
Spark timing has a significant effect on engine combustion and emissions. Regarding engine
efficiency and thus fuel economy, greater expansion work can be produced on the piston as
the spark is advanced and the pressure is consequently increased at TDC. However,
advancing the spark also increases the compression work of the engine as the in-cylinder
11
pressure is increased during the compression stroke. Thus, there is a trade-off between the
compression and expansion work in the IC engine to achieve the best engine efficiency, and
this is controlled by the spark timing (Heywood, 1988). An optimum MFB50 phase of 8-
10˚aTDC is required to produce best engine efficiency and thus the best fuel economy (De O.
Carvalho et al., 2012). At low to medium engine loads, the spark timing can be chosen in
order to produce the optimal combustion phasing. However, at high engine loads, knock
limits the spark advance that can be achieved, meaning that more retarded spark timings are
required (Checkel and Dale, 1989). Emissions are also significantly affected by the engine
spark timing (Heywood, 1988); something that will be discussed in detail in a later section.
The phenomenon of knock has existed for the majority of the history of IC engines, with the
most widely publicised solution to the problem being high octane fuels (Checkel and Dale,
1989, Kalghatgi, 2013). These increase the knock tolerance of the engine which in turn
allows the spark timing to be advanced for fuel economy and torque benefits. However, they
also increase the cost of the fuel because high octane fuel fractions are more expensive than
those with a low octane number (Checkel and Dale, 1989). The fuel economy benefits also
reduce the engine CO2 output, which along with fuel consumption is at the forefront of the
public perception (Gov UK, 2011, Kalghatgi, 2013). Therefore, it is anticipated that demand
for higher octane fuels will only increase in the future (Green Car Congress, 2013, Kalghatgi,
Knock itself is a shock wave generated by the auto-ignition of the fuel-air mixture in the end-
zone of the combustion chamber once it has exceeded a critical temperature of approximately
625-725˚C (Kalghatgi, 2013). This most commonly occurs when the spark timing is
12
advanced to an extent that the pressure and temperature in the engine cylinder are increased
to a critical level where the flame front generated from the spark discharge radiates
significant heat to the fuel-air mixture in the end-zone, causing its pressure and temperature
to increase through compression (Heywood, 1988). This occurs because the density of the
burned mixture increases by approximately a factor of four causing not only the unburned
mixture to be compressed but also compressing those parts of the charge that have already
burned and displacing them back towards the spark plug (Heywood, 1988). This causes the
temperature of the fuel-air mixture in the end-zone to rise beyond 625-725˚C resulting in
The auto-ignition produces two or more flame fronts in the combustion chamber which
rapidly radiate heat to the remaining unburned fuel-air mixture, causing it to combust more
rapidly than in normal combustion (Rothe et al., 2006). When its resulting shock wave makes
contact with the piston face, it produces structural damage rather than exerting a pressure to
push it (Oppenheim, 2004). The high pressure and temperature rise rates produced by the
knock produce very high mechanical and thermal loads in the pistons, bearings, cylinder head
and the cylinder head gasket, causing damage and even failure to these components (Amann
and Alger, 2012, Robert Bosch GmbH, 2006). The flame front of this shock wave can
propagate at a rate of 10 to 100 times that of normal combustion and it resonates on the
combustion chamber walls producing the audible “ping” which is characteristic of engine
Whether knock occurs or not reflects the outcome of a race between the advancing flame
front and the pre-combustion reactions in the unburned end gas. Knock will not occur if the
flame front consumes the end gas before these reactions have time to cause the fuel-air
13
mixture to auto-ignite (Heywood, 1988). However, if these reactions are given sufficient
time, they will produce auto-ignition before the flame front arrives (Heywood, 1988).
There are several factors which affect the tendency of an SI engine to knock. The most
common is substantial spark advance which produces higher in-cylinder temperatures which
leads to higher temperatures in the end-zone of the combustion chamber, resulting in ever
greater levels of knock as the spark is continually advanced (Ulsoy et al., 2012). As the
torque and load demand from the engine increase, the charge density is increased which leads
to higher temperatures during the compression stroke, causing the engine knock tendency to
increase (Robert Bosch GmbH, 2006). By definition, lower octane number fuels will knock
more than those with a higher octane rating, thus, knock tendency increases with their use.
Increasing the compression ratio increases knock tendency because the pressures and
temperatures within the fuel-air charge are increased during the compression stroke (Robert
Bosch GmbH, 2006). Deposits in the combustion chamber also increase the effective
compression ratio, increasing knock tendency (Rajput, 2005, Robert Bosch GmbH, 2006).
Ineffective cooling within the engine can lead to hot spots forming within the combustion
chamber and it can lead to overall temperature increases; both of which increases knock
tendency (Rajput, 2005). Poor turbulence and swirl characteristics of an engine, produced by
the relevant engine geometries can increase knock tendency (Robert Bosch GmbH, 2006).
The use of homogenous fuel injection increases knock tendency because the mixture in the
end-zone of the combustion chamber is ignitable whereas with stratified injection, only the
mixture around the spark plug is rich enough to ignite (Robert Bosch GmbH, 2006). NO
introduced into the intake gas by EGR increases knock tendency (Hoffmeyer et al., 2009,
Lewis et al., 2014, Roth et al., 2010, Takaki et al., 2014, Yuan et al., 2015), but this is
14
overcome by the higher heat capacity of the exhaust gas compared to fresh air (Francqueville
and Michel, 2014), resulting in knock tendency reducing overall. Finally, the use of PFI
injection increases knock tendency in comparison to DI because it does not benefit from the
charge cooling effect that occurs with DI (Alger et al., 2000, Harada et al., 1997).
decrease the engine capacity while maintaining or improving power output, in order to
improve fuel economy and reduce emissions, through reduced pumping losses and other
efficiency gains (Cairns et al., 2006, Castagné et al., 2003, Lecointe and Monnier, 2003).
Technologies such as turbocharging and DI are used in order to achieve these aims (Castagné
et al., 2003, Lecointe and Monnier, 2003, Song et al., 2014). In addition, techniques such as
EGR, swirl and tumble intake airflows and increased compression ratio are being used to
improve the fuel economy and thus reduce the CO2 output of modern engines. Therefore,
EGR, swirl and tumble intake airflows and compression ratio increase have been researched
further in this investigation to help develop the knowledge of engine researchers in these
2.3.5 EGR
EGR, both internal and external, as well as lean mixtures, can be used to reduce the power
density and thus reduce the energy deposition rate of the mixture to improve engine knock
resistance (Francqueville and Michel, 2014, Galloni et al., 2013, Nishida et al., 1988),
(Galloni et al., 2013). This however reduces the maximum engine load and torque, as well as
the maximum power density, meaning that it can only be utilized at part-load conditions. It
can most effectively be used in fast-burn engine designs such as the widely used four-valve
15
pentroof combustion chamber; the design produces significant tumble which leads to a rapid
burn (Stone, 1999, Urushihara et al., 1995). It is most effective in these designs because EGR
reduces the laminar flame speed of the fuel-air mixture, thus, a fast-burn balances the EGR
well (Endres et al., 1992, Heywood, 1988). This allows EGR to be utilised to reduce NOx
emissions, which is particularly important for stratified combustion at low loads, or it allows
it to be utilised to improve part-load fuel economy (Davis and Borgnakke, 1982, Heywood,
1988). However, EGR potentially increases particulate and HC emissions, thus, this limits its
use in modern DISI engines (Alger et al., 2009a, Diana et al., 1996, Wei et al., 2012, Zhong
et al., 2013).
Part-load fuel economy is improved because the throttle has to be more opened to achieve a
particular load, due to the EGR gases replacing some of the fresh air volume in the
combustion chamber (Stone, 1999). This reduces the pumping work required of the engine,
thus improving fuel economy (Abd-Alla, 2002, Grandin et al., 1998, Robert Bosch GmbH,
2006, Wei et al., 2012). This is important at high loads because stratification for the purpose
of reducing pumping losses is not possible. Thermal efficiency is also improved with EGR
addition due to reduced combustion temperatures (Ratnak et al., 2015, Siokos et al., 2015),
contributing to fuel economy improvements. As mentioned, the spark timing can be advanced
due to the improved knock resistance with EGR addition, allowing the optimum 8-10˚aTDC
MFB50 phasing to be achieved, also improving the fuel economy (De O. Carvalho et al.,
2012). EGTs are also reduced with EGR which can be useful for protecting the inlet turbine
of turbocharged engines (Alger et al., 2009a). However, the EGR system must be sufficiently
robust to withstand the deposits that accumulate in the system due to the lower EGTs
16
Another drawback of EGR is that COVIMEP is increased, and eventually some cycles become
sufficiently slow burning with high EGR ratios that combustion is not completed by the time
the exhaust valve opens, eventually causing engine misfire (Matsushita et al., 1985, Siokos et
al., 2015, Wei et al., 2012). Thus, it is necessary to advance the spark in order to ensure that
the combustion is sufficiently completed before the exhaust valve opens. Depending on the
engine operating condition, the optimum 8-10˚aTDC MFB50 phasing can either be achieved
at a certain spark timing and EGR addition, or the combustion can degrade so much that
despite advancing the spark timing, the combustion duration is too long to achieve a
sufficiently advanced MFB50 combustion phasing, resulting in partial burns and misfire (Pan
et al., 2014). Therefore, there is a maximum rate of EGR that can be achieved while keeping
is reduced at high EGR levels due to the increased combustion duration (Siokos et al., 2015,
Typically higher engine loads are more tolerant to EGR addition but again, there is a trade-off
because as mentioned, the diluting effects of EGR reduce the maximum load that the engine
can achieve, and no EGR addition can be achieved at full-load without using an external
pump due to the equivalent intake and exhaust pressures. Thus, the maximum EGR addition
Cooled EGR has been shown to be more effective in suppressing engine knock than hot EGR
due to reduced temperatures in the end-zone of the combustion chamber (Alger et al., 2009a),
while EGR after a TWC (catalysed EGR) further reduces knocking tendency because NO
present in EGR gases before a TWC induces knock (Hoffmeyer et al., 2009, Lewis et al.,
2014, Parsons et al., 2015). The reduced knock allows the spark timing to be advanced
17
resulting in fuel economy improvements. It also allows the compression ratio to be increased,
also resulting in fuel economy improvements (Su et al., 2014). Fuel economy is also
improved as a result of the lower in-cylinder temperatures and thus improved thermal and
volumetric efficiencies achieved by cooling the EGR gases (Su et al., 2014, Wei et al., 2012).
However, the maximum EGR addition achievable with cooled EGR may be limited because
of water droplet formation (Siokos et al., 2015) and it has been observed to increase COV IMEP
and HC emissions (Wei et al., 2012). Despite the work that has been done, much more
investigation needs to be conducted into the effect of different EGR types (hot EGR, cooled
EGR, hot EGR after TWC, cooled EGR after TWC) on engine combustion and emissions.
Most commonly, EGR is used in its homogenous form where the intake air is well-mixed
with the EGR gases. However, some researchers have investigated stratified EGR to help
overcome the reduction of burning speed, the HC emission increase and the increase in
COVIMEP (Wei et al., 2012). This is where the fresh air and EGR gases are separated in the
combustion chamber, which improves flame propagation at the spark plug region (Wei et al.,
2012). However, there are significant challenges in implementing stratified EGR. Due to the
flow structure in the combustion chamber being extremely complex, it is very difficult to
achieve complete separation of the fresh air and EGR gases, and it is challenging to maintain
the stratification throughout the intake and compression strokes prior to ignition (Wei et al.,
2012). Therefore, this is why stratified EGR is not pursued in this investigation.
induction and compression strokes of the engine cycle, along with the geometry of the
combustion chamber (Dunn-Rankin, 2008). Swirl is defined as organized rotation of the fuel-
18
air charge about the cylinder axis, while tumble is defined as organized rotation of the fuel-air
charge around an axis perpendicular to that of the cylinder axis, at half the height of the
combustion chamber (Matsushita et al., 1985, Stone and Ladommatos, 1992). While the swirl
motion will undergo some decay during the engine cycle due to friction, intake swirl usually
persists through the compression and combustion, and it is converted into general turbulence
Swirl is typically used to improve fuel air mixing (improved mixture homogeneity) and to
increase the speed of the combustion process, particularly with air-guided fuel injection in
order to transport the fuel-air mixture towards the spark plug (Agarwal et al., 2014,
Nagayama et al., 1977). For pre-mixed combustion in SI engines, the effect of increased in-
cylinder turbulence breaks-up or wrinkles the flame front, which has the effect of increasing
the area of the flame front increasing the combustion speed (Lee, 2007, Stone, 1999).
Increasing the in-cylinder turbulence allows leaner mixtures to be burned which are less
prone to knock, due to the reduced combustion temperatures; improving fuel economy and
affecting engine emissions (Nagayama et al., 1977, Stone, 1999). PM emissions, as well as
NOx and HC emissions will be affected, even with stoichiometric conditions (Kim and Kim,
1995, Mikulec et al., 1988, Urushihara et al., 1995). Indeed, (Mehta et al., 2001) reported that
the use of swirl valves resulted in significant improvements in PM mass as compared to low
speed is increased by creating consistent in-cylinder flow conditions (Zeng et al., 2015).
turbulence due to the faster combustion; the fuel-air charge in the end-zone of the combustion
However, it must be noted that a faster flame will also increase in-cylinder pressures which
19
will make auto-ignition more likely. In general, faster burning cycles are more prone to knock
(Kalghatgi, 2013, Leppard, 1982). Furthermore, swirl allows EGR, as discussed previously,
A disadvantage of generating engine swirl and tumble intake airflows is that the volumetric
efficiency (VE) is reduced which can cause the fuel economy to worsen (Arcoumanis and
Kamimoto, 2009). This also limits the EGR addition that can be achieved, since there is a
smaller pressure difference between the exhaust and intake. Although the use of lean
mixtures will reduce the efficiency of a TWC causing emissions to increase, the emissions
improvements produced by the lean burn combustion may be sufficient to offset this loss
(Stone, 1999). Finally, excessive tumble can lead to an increase in HC emissions and an
Using swirl and tumble along with EGR addition allows the engine to operate under learn
burn conditions in order to improve fuel economy and NO x emissions, while maintaining HC
emissions at the same level (Endres et al., 1992). In addition, high tumble airflows can be
used to achieve high EGR additions of up to 30-40% which can be used to improve fuel
economy (Ikeya et al., 2015, Takahashi et al., 2015, Zhang et al., 2014a). High swirl airflows
could also be used for the same purpose. COVIMEP can also be reduced with a combination of
tumble and EGR addition as compared to using EGR alone, due to enhanced turbulent flow in
the cylinder (Zhang et al., 2014a). In particular, because EGR allows the spark timing to be
advanced, it allows the point of ignition to occur earlier in the engine cycle where the
turbulence is stronger. This increases the laminar flame speed allowing the combustion
20
2.3.7 Compression Ratio
Increased compression ratios can lead to significant fuel economy improvements (Gumbleton
et al., 1976, Heywood, 1988). In an engine in which the compression ratio was raised from
9.7:1 to 14.6:1, an efficiency gain of up to 15% was obtained with larger gains at part throttle,
due to the improved thermal efficiency (Stone, 1999). The EGT is decreased when
compression ratio is increased, due to more efficient conversion of the fuel energy into work
However, there are also disadvantages to increased compression ratios. The fuel
consumption can worsen under low-speed, high-load conditions at high compression ratios
due to spark retardation caused by heavy knocking (Okamoto et al., 2003). NOx emissions
increase due to higher in-cylinder temperatures. Large squish areas and poor surface-to-
volume ratios increase the hydrocarbon emissions (Stone, 1999). This is in addition to the
increased hydrocarbons resulting from the higher in-cylinder pressures which cause more
HCs to be stored in the piston crevices (Heywood, 1988). When released during the
expansion stroke these do not burn fully, thus contributing to HC emission increases with
increased compression ratios (Heywood, 1988, Zhang et al., 2014b). They also result from
the lower EGTs which reduce the rate of HC oxidation (Heywood, 1988). Finally,
combustion deposits become more significant as the compression ratio is increased because
the ratio of their volume compared to the volume of the combustion chamber increases (Liiva
Compression ratio increase with EGR addition has been shown to significantly improve fuel
economy while reducing the sum of the NOx and HC emissions (Diana et al., 1996). The
21
combination of the two has also been shown to produce a relatively low COV IMEP due to
increased turbulence intensity and reduced cyclic variations (Pan et al., 2014).
order to form new bonds with oxygen atoms when the fuel is burned in air. This process
produces a significant amount of energy which is converted into work on the piston which
2.4.1 Gasoline
The size and geometry of the hydrocarbon molecule has a significant effect on its chemical
properties, providing a wide-range of potential fuels that can be utilized in IC engines (Stone,
1999). Indeed, gasoline is a mixture of many different sized alkane and alkene hydrocarbon
molecules with varying chain lengths and isomerisation. Hydrocarbons can also form
aromatic and cyclo-alkane ring structures such as benzene, which are also present in gasoline
Overall, the two most important characteristics of gasoline are its volatility and octane
number (Stone, 1999). Gasoline fuel is required to vaporize well within a given temperature
range that is season dependent in most countries so that it can produce an ignitable fuel-air
mixture when injected into the combustion chamber. If it vaporizes too easily then it can
result in problems such as vapour lock and excessive evaporative emissions (Srinivasan,
2001). However, if it is not volatile enough it will be difficult to produce an ignitable fuel-air
mixture in cold conditions, leading to poor cold startability (Kalghatgi, 2013, Robert Bosch
GmbH, 2006). If the fuel does not vaporize well enough it can also pass into the engine oil on
the combustion chamber walls, causing it to become diluted (Kalghatgi, 2013, Robert Bosch
22
GmbH, 2006). Finally, poorly vaporized fuel can lead to excessive deposit formation (Liiva
et al., 1992). Fuel octane number limits the spark timing advance that can be achieved with
the engine, and thus the fuel economy and torque, because of engine knock, as discussed.
Thus, fuel octane number requirements have increased and will continue to increase in the
coming years, in order to achieve further fuel economy improvements (Green Car Congress,
2013). The fuel specifications for gasoline fuel in the EU and US are provided in Table 2-1.
Table 2-1 EU (EN 228-2012) and US (ASTM D4814 Rev B-2011) Fuel Specification
(Analiit AA, 2015, Chevron, 2009)
Parameter EU US
Density (kg/m³) 720-775 (at 15°C) -
1
RON 95 -
MON 85 -
Distillation Properties
T10 (°C) - ≤70
T50 (°C) - ≤77-121
T90 (°C) - ≤190
End of distillation point (°C) - ≤205
Evaporation Properties
Percentage evaporated at 70˚C
20-48 -
summer (volume %)2
Percentage evaporated at 70˚C
22-50 -
Winter (volume %)2
Percentage evaporated at 100˚C
46-71 -
(volume %)
Percentage evaporated at 150˚C
≥75 -
(volume %)
Oxygenates
Aromatics (m/m) ≤35% ≤20.9%
Olefins (%m/m) ≤18% ≤11.9%
Sulphur (ppm) ≤10 ≤80
1
Member States may decide to continue to permit the placing on the market of unleaded regular grade petrol with a
minimum motor octane number (MON) of 81 and a minimum research octane number (RON) of 91.
2
Depends on volatility class of region, country and season
The octane number scale is used to define the octane number for a given fuel; n-heptane is
assigned an octane number of 0 and isooctane is assigned an octane number of 100. The
octane number of a given fuel is thus defined as the mixture of n-heptane and isooctane
which produces the equivalent knocking tendency of the given fuel. For example, if the fuel
23
has the same knocking tendency as a fuel mixture of 95% octane and 5% heptane, then it is
given an octane number of 95 (Burton et al., 2000). For octane numbers above 100, a
tetraethyl lead knock additive is added to the isooctane fuel in order to allow the octane
The two most widely used tests for defining the knock tendency of a given fuel are the
Research Method (ASTM D-2699) and the Motor Method (ASTM D-2700), both of which
The test for MON differs from that used for the RON test by using pre-heated mixtures,
higher engine speeds and variable ignition timing. Thus, the fuel being examined is subjected
to more stringent thermal demands, which is why MON figures are typically lower than those
for RON (Bradley et al., 2004, Robert Bosch GmbH, 2006). The engine itself was developed
CFR engine (Bock et al., 2013). The CFR engine is a robust four-stroke overhead valve
engine with a compression ratio that can be varied from 3 to 30 while operating (Heywood,
1988). The operating conditions for the two tests are shown in Table 2-2.
Table 2-2 Operating conditions for research and motor octane number test methods
(Bradley, 2004, Heywood, 1988)
24
Overall, the test conditions are selected to represent the engine operating range where knock
is most severe. The air-fuel ratio is adjusted to achieve the most severe knock, then the
However, there are some limitations of the standard tests conducted in the CFR engine; both
RON and MON are evaluated at low engine speeds and the air/fuel ratio is adjusted to
produce the maximum knock (Stone, 1999). All practical gasolines are mixtures of aromatics,
olefins, naphthenes, oxygenates and paraffins which have auto-ignition chemistries different
from those of n-heptane and isooctane (Heywood, 1988). Compared to primary reference
fuels (PRF’s), real fuels become more resistant to knock as pressure is increased for the same
temperature, thus, the MON method of increasing the intake temperature is limited.
Furthermore, because alcohols burn more rapidly than gasoline, their ignition timings in the
RON and MON tests are over-advanced, leading to an underestimate of their knock
performance (Heywood, 1988). Therefore, engines can behave rather differently during their
daily use from what these standard tests would otherwise suggest. Due to these limitations, an
average of the RON and MON values is typically taken in order to produce an antiknock
index which offers an improved characterization of the antiknock properties of the fuel
Fuel volatility is significantly affected by the chemical structure of the fuel, with shorter
carbon chains having a higher volatility than longer chains, and branched carbon chains
having a higher volatility than straight chains with the same number of carbon atoms (Burton
et al., 2000). The chemical structure of the fuel also significantly affects its RON and MON.
Fuel octane number decreases with increasing chain length, it increases with increasing
numbers of side chains for the same number of carbon atoms and it increases with ring
25
structures such as those found in cyclo-alkanes and aromatics (Kalghatgi, 2013). This has
significant effects on achieving ever higher octane number fuels, because of concerns over
(although many are still made from crude-oil stocks), and along with other oxygenated
molecules, they are important components of modern day gasoline fuel (Hamilton et al.,
2008, Heywood, 1988). The main drawback of oxygenated fuels is their lower gravimetric
calorific values as compared to pure hydrocarbons, due to their partially oxygenated nature
(Daniel et al., 2012a, Kalghatgi, 2013). This typically results in higher engine-out CO2
emissions (Al-Hasan, 2003, Hsieh et al., 2002, Masum et al., 2014, Varol et al., 2014)
because more fuel needs to be burned in order to achieve the same power output. The reduced
power output can be overcome with compression ratio increases however, with 40-60%vol
ethanol in gasoline blends being shown to have a similar performance to gasoline (Goswami
et al., 2015), thus reducing the CO2 increases. Furthermore, alcohols such as ethanol have a
high octane number and high enthalpy of vaporization which allows the spark timing to be
advanced as compared to regular gasoline fuel, enabling the fuel consumption penalty to be
reduced and torque to be improved (Al-Hasan, 2003, Deng et al., 2013a, Irimescu et al.,
2015, Poitras et al., 2015). The high enthalpy of vaporization also reduces the charge
temperature which improves VE and the lower air/fuel ratios mean that more energy is
released per kilogram of stoichiometric mixture burned; both of these lead to overall
Oxygenated molecules such as ethanol and butanol burn more quickly than gasoline, leading
to enhanced combustion stability and improved engine efficiency (Deng et al., 2013b, Fu et
26
al., 2014, Masum et al., 2013, Turner et al., 2011). CO emissions are typically reduced due to
the greater oxidation resulting from the fuel oxygen content (Al-Hasan, 2003, He et al., 2003,
Karavalakis et al., 2014, Varol et al., 2014, Yejian et al., 2007). Reduced combustion
temperatures with oxygenated fuels reduce the NOx emissions and EGT (Elfasakhany, 2014)
and their oxygenated nature increases the rate of oxidation of PM and HC emissions (Chen et
al., 2015, Goswami et al., 2015, Gu et al., 2012, Irimescu et al., 2015, Mattimaricq, 2012,
Wang et al., 2014a, Zhang et al., 2015). Furthermore, (Wang et al., 2014c) observed that soot
produced from the combustion of oxygenated fuels has lower activation energies than that
produced by gasoline, resulting in increased oxidation rates in the exhaust. Some researchers
have reported increased NOx emissions with oxygenated fuels due to their increased fuel
oxygen content (Bayraktar, 2005, Feng et al., 2013, Graham et al., 2008, Keskin et al., 2011,
Masum et al., 2014) which is important to note. However, this may result from the air/fuel
ratio in their engines not being changed from that of gasoline when the oxygenated fuel was
On the other hand, despite the improved knock properties resulting from the higher heat of
vaporization, the cold start performance can worsen because the fuel becomes harder to
vaporize (Aleiferis et al., 2013). The higher heat of vaporization combined with the typically
reduced vapour pressures and boiling points of oxygenated fuels in comparison to gasoline
means that they have to be injected earlier in the engine cycle to form a homogenous mixture,
reducing flexibility in engine calibration. Oxygenated fuels can absorb moisture which can
lead to corrosion and fuel separation problems (Hu et al., 2012, Stone, 1999). Alcohols can
also damage materials used in the fuel injection equipment and they may cause elastomer
27
Hydrous ethanol provides benefits over conventional anhydrous ethanol such as lower cost
due to the simpler manufacturing process (final ethanol dehydration step is removed) and
reduced NOx emissions resulting from lower peak in-cylinder temperatures (its water content
absorbs heat) (Masum et al., 2013). However, problems arise when the hydrous ethanol is
blended into gasoline fuel because of the immiscibility of water and gasoline, limiting the
amount of water that can be used in such mixtures (Stone et al., 2012).
1-butanol while not widely used in 1-butanol-gasoline blends has the potential to be used in
the future with similar blend ratios as ethanol-gasoline blending; therefore, 1-butanol has
been researched in this investigation (Fountain, 2012). This is because its energy content is
higher than ethanol, butanol is less prone to water contamination and it is less corrosive than
ethanol (Szwaja and Naber, 2010). Therefore, this will allow it to be distributed in the same
infrastructure that is currently used for gasoline, significantly reducing the infrastructure costs
as compared to ethanol (Szwaja and Naber, 2010). It must be noted however that 1-butanol
has been observed to have poorer cold startability as compared to ethanol, because it is harder
to ignite at cold conditions (20˚C) (Serras-Pereira et al., 2013), possibly limiting the blend
ratio of 1-butanol in gasoline. Its production costs are also higher (Masum et al., 2013)
providing a further drawback. Although 2-butanol (iso-butanol) has potential advantages over
1-butanol (higher octane number of 101 compared to 96 for 1-butanol, lower boiling point of
99.5˚C compared to 117.7˚C for 1-butanol resulting in better cold startability) (Liu et al.,
2013), it was not investigated because of the greater availability of 1-butanol in the
marketplace.
Ethanol is used on a wide scale only in its blended forms of up to 10%vol and in the near
future this trend is likely to grow with ethanol-gasoline blends between 20-40%vol (Hamilton
28
et al., 2008, Stein et al., 2013, Wallner et al., 2013). This is being driven by a 10% minimum
target for the use of alternative fuels in transportation in the EU legislated by Directive
2009/28/EC in 2009, and by the Renewable Fuel Standard in the Energy Independence and
Security Act of 2007 in the USA (Daniel, 2012d, Directive 2009/28/EC, 2009, Splitter and
Szybist, 2014). Therefore, ethanol fuel blending in gasoline has also been researched in this
investigation. In order to improve the usefulness of the investigation, the blended form is
tested which not only allows the effect of ethanol addition to gasoline on DISI engine
combustion and emissions to be quantified; it also allows the precise effects of one of the
quantified.
2.5.1 PM Emissions
Solid particles created as a result of incomplete combustion are known as particulates which
consist of chains of carbon particles, i.e. soot, onto which uncombusted or partially
combusted hydrocarbons form deposits (Robert Bosch GmbH, 2006). According to (Zhang et
al., 2014b), the nucleation mode particles mainly result from droplets formed by hydrocarbon
particles are mainly composed of carbonaceous agglomerates formed in local rich-fuel zones
polycyclic aromatic hydrocarbon (PAHs) and accumulation mode particles (Zhang et al.,
2014b).
The Euro VI (2014) emission legislation which has for the first time limited particulate
number (PN) for DISI engines, in addition to the mass limit already imposed by Euro V
29
(2009), poses a significant challenge for engine researchers; thus, it is an important area of
research (Hedge et al., 2011). The PN emissions are higher than those produced by a diesel
engine with a particulate filter; therefore, this is one reason why the legislation has become
Due to the longer period of time they have been regulated for, PM mass emissions have
fine PM particles, in particular, have a negative impact on human health (Anderson et al.,
2012b, Hull et al., 2005, United States Environmental Protection Agency, 2014). Therefore,
there is a greater need to reduce PN which is why it is studied in detail in this investigation,
rather than PM mass. This is despite the PN being less consistent than PM mass (Eastwood,
2008). The smallest particles are at the limit of detection of PM measurement instruments,
causing the numbers of smaller particles recorded to vary more significantly than larger
particles which are more easily detectable (Eastwood, 2008). These smaller sized particles are
often the most numerous in the exhaust emission resulting in the more significant variation in
PN (Eastwood, 2008).
stratified combustion (late fuel injection) due to less nuclei adsorption onto accumulation
mode particles occurring, because a lower number of accumulation mode particles are
result of reduced piston wetting and improved mixture formation (Eastwood, 2008). For this
reason, homogenous combustion has been studied in this investigation (start of injection
(SOI) =280˚bTDC) at medium to high engine loads (5.5-8.5 bar IMEP) where homogenous
30
Overall, the regulations are dictating the current and future direction of engine development
with growing emphasis being placed upon the reduction of PM mass and number based
emissions from DISI engines (Delphi, 2012). PFI produces less PM than DI because there is
typically more time for fuel-air mixing before combustion (Alger et al., 2000, Zhang et al.,
2014b) and there is no wall wetting and fuel impingement onto the piston crown as there is
with DI. Thus, the PN emissions of different vehicle types can be summarized as:
Particle formation in DISI engines can be classified into two stages. The first stage is the
formation of the particles; the first condensed phase material arises from the fuel molecules
via their oxidation and/or pyrolysis products which typically include various unsaturated
these result in the production of nucleation particles. The second stage of particle growth,
which includes both surface growth and aggregation, is where the bulk of the solid-phase
material is generated. This involves the bonding of gas-phase species to the surface of the
particles and their incorporation into the particulate phase (Heywood, 1988). Gaseous
particles also adsorb onto the particles in the engine exhaust (Heywood, 1988).
Overall, the formation mechanism of particulates in DISI engines is much the same as in
Diesel engines in the sense that it is highly dependent on local fuel equivalence ratio and
charge temperature (Piock et al., 2011). The greatest formation rates are thus typically found
in areas of the combustion chamber with a high local fuel equivalence ratio (rich) and a low
local charge temperature. However, it is important to note that lower charge temperatures can
also decrease PM formation through reducing the numbers of primary carbon particles
formed by the thermal pyrolysis and dehydrogenation reaction of fuel vapour/droplets (Zhang
et al., 2014b), thus complicating PM behaviour. Oxidation of the particulates occurs post-
31
combustion in the expansion and exhaust strokes, and in the exhaust system of the engine,
converting them to CO and CO2 as well as less complex particles (Heywood, 1988, Piock et
al., 2011).
EGR is generally considered to have a negative effect on engine particulates, however, not
many research studies have been published investigating this issue (Alger et al., 2009a,
Zhong et al., 2013). From the limited research available, it can be concluded that EGR
reduces the need for fuel enrichment to reduce EGTs in turbocharged engines, causing PM
mass and number to reduce significantly (Alger et al., 2010). However, much more research
oxidize once it is released from the engine and it can go on to react with unburned non-
(Stone, 1999). Therefore, there is little benefit in measuring just NO or NO2 output from an
engine; rather it is more useful to measure NOx. NO forms in the high temperature burned
gas, where there is sufficient temperature for nitrogen in the air to oxidise, which typically
occurs at temperatures above 1525˚C (Heywood, 1988, Nieman et al., 2012). Pure NO2 is a
toxic reddish-brown gas with strong odour which can cause irritation of mucous membranes
carcinogenic when people are exposed to them for a long time (Robert Bosch GmbH, 2006).
In addition, some of their products produced when they are exposed to sunlight are also
32
Emissions of NOx and HC vary between different engines and are affected by ignition timing,
load, EGR ratio, speed and air/fuel ratio (Stone, 1999). Regarding spark timing, typically
advancing the spark timing will increase in-cylinder pressures and temperatures, increasing
the formation of NOx and increasing the oxidation of HCs in the combustion chamber,
resulting in increased NOx and decreased HC. Retarding the spark timing will typically
reduce the in-cylinder pressures and temperatures, causing the NOx emissions to decrease and
the HC emissions to increase. Their real-life behaviour however, is more complicated than
this simplification. More advanced spark timings will lead to poorer mixture preparation,
which alone would lead to higher HC emissions. However, as the spark becomes more
advanced, the time available for post-combustion oxidization is increased, which alone would
lead to lower HC emissions. These factors interact with each other, which can sometimes lead
to unexpected PM emission behaviour. Increasing the engine load will increase the in-
cylinder pressure and temperatures, which will typically increase NO x emissions and reduce
HC emissions are typically increased (Alger et al., 2009a, Davis and Borgnakke, 1982, Zhong
et al., 2013).
Overall, HC emissions in SI engines originate from several sources. Firstly, unburned fuel-air
escapes into the exhaust during the valve overlap period (Stone, 1999). Secondly, misfires or
partial burns in which the mixture fails to ignite or the combustion was too slow or started too
late, respectively, causes HC emissions (Heywood, 1988, Robert Bosch GmbH, 2006).
Thirdly, absorption/desorption from oil films, combustion deposits and crevices occurs as the
pressure is rising and falling within the engine cylinder (Heywood, 1988, Tomita et al.,
1996). Fourthly, a quench layer of unburned and partially burned fuel is left at the
combustion chamber wall when the flame is extinguished as it approaches it, which can be
33
released during the exhaust stroke (Tomita et al., 1996). Finally, poor mixture preparation,
particularly during transients and cold start, leads to the partial oxidization of HCs in the
In this investigation the NO x emission from the engine was studied because as shown in
Table 2-3, the legislation regarding NO x has become stricter from 2009 (Euro V), making
these emissions important to investigate to help further control them. Although the THC
emission regulation has remained consistent since 2005 (Euro IV), they are still useful to
study because of their relationship with the NOx and nucleation mode particulate matter (PM)
emissions (Zhang et al., 2014b). CO emissions are also harmful to human health. The CO
molecule attaches itself to haemoglobin in the bloodstream, preventing oxygen from being
transported by red blood cells, making it harder for the body to circulate oxygen to vital
organs, resulting in nausea, headaches and other health problems. Despite this, CO emissions
were not studied because their regulated numbers have been consistent from 2005 (Euro IV),
thus, it was considered more useful to study other engine combustion and emission
parameters.
subsequent effect on global temperature and climate (Gov UK, 2011). The transportation
growing political drive to reduce CO2 emissions from automobiles (Johnson, 2010, Zaccardi
et al., 2009). While oxygenated fuels typically produce increased engine-out CO2 emissions
as compared to gasoline because of their reduced lower heating value (LHV) which increases
fuel consumption, as discussed, they are typically produced using renewable methods.
Therefore, net CO2 output from the use of these fuels is reduced (Daniel, 2012c et al., Owen
34
and Coley, 1995, Parag and Raghavan, 2009, Turner et al., 2011). Improvements in fuel
consumption also reduce CO2 output from the transportation sector (Zaccardi et al., 2009).
Therefore, oxygenated fuels and fuel consumption have been researched further in this
investigation.
2.5.4 Legislation
The evolution of European Union emission legislation for DISI petrol cars with a mass of less
Table 2-3 European Union emission legislation for DISI petrol cars (Delphi, 2012)
The emissions limit themselves were defined firstly in 1970 (Directive (70/156/EC)) and they
are tested over the New European Drive Cycle (NEDC); a synthetically generated test cycle
consisting of phases of constant cruising speeds and constant rates of acceleration (Robert
Bosch GmbH, 2005). In the near future, they will be tested over a Real Driving Emission
(RDE) test cycle to more accurately measure the pollutant emissions by using real world
driving conditions rather than a synthetically generated test cycle, which bears little relevance
considered to be more stringent than the European emission legislation, with the Federal Test
Procedure (FTP) drive cycle mirroring actual highway driving conditions which are harder to
35
specifically calibrate an engine for. This legislation is summarized in Table 2-4; it is
Table 2-4 United States Federal emission legislation for Light Duty SI Vehicles <80,000
km (Delphi, 2012)
NOx 250 90
mg/km
CO 2100 1050
PM (mass) 50 15
PM (number) #/km - -
There is a certain degree of flexibility with the US legislation because each vehicle
manufacturer must only ensure that the exhaust emissions for its total vehicle fleet do not
exceed the specified average; not every vehicle from the fleet must meet the regulations
(United States Environmental Protection Agency, 2010). Only a certain percentage of the
manufacturer’s fleet must comply with the legislation too, but the percentage is increasing
year-on-year in order to reduce the emissions across the vehicle fleets of manufacturers
For both legislations, vehicles are randomly tested during their lifetime in order to ensure that
their emissions have not become too degraded, thus adding an extra layer of difficulty for
36
2.5.4 TWCs
Despite the ability of TWCs to reduce the emissions from a gasoline vehicle significantly,
produced by the engine itself. This is because emissions increase significantly during cold-
start when the catalyst is not at its optimum operating temperature, causing significant
emissions to be produced by the vehicle (Frennet and Bastin, 1995). The NEDC begins from
when the engine is started so if the emissions produced by the engine alone are not below a
certain limit, the emissions produced during cold-start could potentially cause the vehicle to
fail the emissions test (Isermann, 2014). In addition, when the engine undergoes acceleration,
the air/fuel charge is typically enriched as well as during cold-start, meaning that the TWC
condition (Alger and Mangold, 2009). Finally, there is a strong relationship between the
nucleation mode particles and HC emission (Zhang et al., 2014b); therefore, it is still useful
2.6 Summary
Overall, this literature review discussed the development of modern DISI technology,
particularly in relation to fuel consumption improvement and emissions reduction. The major
areas include discussion of the knocking phenomena, EGR and relevant emission regulations.
Engine downsizing was introduced after which fuel economy improving and thus CO2
reducing techniques of EGR (also used to reduce NOx emissions), swirl and tumble intake
airflows and compression ratio increase were discussed. SI fuels and emission regulations
then follow due to their strong influence, particularly the latter, on the development of IC
engines.
37
The main focus of the review is to provide a background to the fuel economy improving and
thus CO2 reducing techniques which are researched further in this investigation. Compression
ratio increase helps to accomplish the objectives of engine researchers to increase engine
efficiency and thus reduce fuel consumption, producing subsequent CO2 emissions benefits.
However, knock, which has been discussed in detail, is a limiting factor regarding the
maximum compression ratio that can be achieved, limiting engine efficiency. Therefore, EGR
is used to enable higher compression ratios to be achieved while reducing the engine NOx
reduced compared to their equivalent NA engines (Su et al., 2014), in order to achieve a
higher IMEP, this reduction can be reduced with EGR addition. Swirl and tumble intake
airflows are used to improve fuel-air mixing in the combustion chamber for improved
emissions, as well as increasing the combustion speed, to enable higher EGR levels and thus
Spark ignition fuels were discussed in detail due to their significant effect on the engine
emissions, with oxygenated fuels explored because of their potential to reduce net CO 2 output
from the transportation sector, if produced by renewable means. The relevant emissions
regulations have been discussed at length due to their significant influence on the
development of IC engines and fuels, with the new Euro VI emission regulation, in particular,
Ultimately, this literature review introduced the main motivation of this thesis which is to
investigate the effect of the fuel economy improving and thus CO 2 emission reducing
techniques of EGR (also used to reduce NOx), swirl and tumble intake airflows, increased
compression ratio, as well as the effect of oxygenated fuel components of 1-butanol and
38
Chapter 3
The aim of this chapter is to present the experimental test facilities and the data acquisition
3.1 Introduction
This chapter describes all of the test equipment used throughout this study. This includes the
details of the single cylinder engine hardware and instrumentation, as well as the details of
the emissions equipment. The engine control methodology, data acquisition systems and
The engine test facilities were created and developed by Jaguar Land Rover and the previous
research students at the University in their PhD studies. During the investigation, the author
was responsible for the development and maintenance of the single-cylinder thermal engine
test facility. Developing the engine test cell facility involved improving the fuel and variable
valve timing systems. Regarding maintenance of the facilities, this included changing minor
components such as fuel pumps, water pumps, water heaters, etc., when they failed, and three
engine rebuilds.
Overall, this chapter provides an overview of the current test facility. Further detailed
descriptions can be found in (Wang, 2014b), (Daniel, 2012d), (Turner 2010) and (Luszcz,
2009).
39
3.2 Overview of Single-Cylinder Engine Test Cell
The schematic of the experimental test facility is shown in Figure 3-1.
Figure 3-1 Schematic of single-cylinder engine test facility and instrumentation setup
The figure shows the 4-stroke single-cylinder spray-guided DISI engine used in this
investigation. The cylinder head is a single-cylinder version of that used in the Jaguar AJ133
(2010) 5.0 litre V8 production engine. This was attached to a balancing unit based on a
Ricardo Hydra engine which was redesigned by the Internal Combustion Engines (ICE)
group at the University of Oxford. The engine cylinder head provides the combustion
chamber with 4-valves and spray guided DI; both of which are found in modern DISI
engines.
The modified single cylinder research engine was not designed to be very resistant to knock
(high compression ratio of 11.5, large bore diameter of 90 mm). In addition, 95 RON
gasoline was used throughout this research which is of a lower octane number than the 98
40
RON gasoline specified for the engine. Therefore, engine knock occurred at loads of 6.0 bar
IMEP and above, which is somewhat lower than what can be expected with the state of the
art aggressively downsized engines of modern cars on sale today. Furthermore, audible knock
was observed to start occurring with 97 RON gasoline fuel at engine loads between 4.5 and
6.0 bar IMEP by previous research students using this research engine (Wang, 2014b, Daniel,
2012d). Therefore, the occurrence of knock at loads of 6.0 bar IMEP and above with 95 RON
gasoline fuel is consistent with these previous investigations. This allowed the effect of EGR
on knock to be studied at reduced loads to prevent any significant engine damage. The engine
also features a variable valve timing (VVT) system which will be discussed in detail in a later
section.
A full specification of the single-cylinder engine is shown in Table 3-1. The overall design of
the engine provides it with an over-square configuration resulting in a high compression ratio
Parameter
The balancing unit was attached to a test bed and the ancillary equipment used to operate the
engine and to record the data was attached onto the engine and balancing unit. This will be
41
Before entering the engine, the intake air flow was filtered after which it passed through the
volumetric airflow (VAF) meter, as shown in Figure 3-2, which recorded the rate of air flow
through the engine. It then passed through a 100 litre damper tank which reduced the pressure
fluctuations through the VAF after which it passed through the throttle and the intake plenum
chamber; the chamber reduced the pressure fluctuations after the throttle. The air then went
into the engine before mixing with and undergoing combustion with the fuel, after which the
combustion products were ejected into the exhaust system. The exhaust gases passed through
the exhaust plenum chamber which reduced the pressure fluctuations before being analysed
by the Lambda meter to measure the air/fuel ratio. The gases were then analysed by the
particulate and gaseous emission measurement systems, after which they were pumped out to
the atmosphere via the engine laboratory ventilation system. An EGR line was connected just
after the exhaust plenum chamber to the intake plenum chamber to achieve external EGR. To
achieve swirl and tumble intake airflows in the engine, swirl and tumble inducing baffle
plates were fitted between the intake runner and cylinder head. These systems are discussed
Mawdsley’s LTD. It used 450 volt electric power and it could maintain a constant speed or
constant engine load with motoring or firing conditions. The speed was manually controlled
by a dial in the engine control room and it was displayed on a digital indicator visible from
and exhaust plenum chambers were used to reduce the pressure pulsations through the intake
42
and exhaust systems in order to stabilize the airflow through the engine. Based on the work of
a previous student (Daniel, 2012d), a 100 litre intake damper was used in order to supply the
engine with stable airflow at wide open throttle (WOT). Smaller plenums wee required just
before the intake and just after the exhaust because the 100 litre intake damper already
effectively reduced the pulsations in the intake and exhaust systems; they were required only
test facility and are shown schematically in Figure 3-2. Each circuit was pumped and heated
individually through water and oil pumps and heaters, respectively. In addition, a large water-
water heat exchanger was used to transfer heat between the warm engine coolant and cold
mains water and another smaller water-oil heat exchanger was used to transfer heat between
the water and oil; both heat exchangers were manufactured by Bowman Ltd. Overall, the
pumping, heating and heat exchange allowed the temperature of the water and oil circuits to
be maintained within a reasonable range. During the experiments, the water and oil
integral differential (PID) controller. Once the temperature exceeded the set limit on the PID
controller, it sent a signal which, via a relay, opened a solenoid valve, allowing cold mains
water into the water-water heat exchanger to cool the engine coolant. This then cooled the
43
Figure 3-2 Schematic of water cooling and oil lubrication circuits
Tap water was used to cool the engine because the engine does not have any specialist
cooling requirements. 10W40 oil was used to lubricate the engine, as recommended by
Jaguar Land Rover. The water and oil levels were monitored and regularly topped up to
ensure there was enough water and oil available to meet the cooling and lubricant
requirements of the engine when motoring and firing. The oil was regularly replaced because
EGR led to degradation of the oil due to an increased number of particulates condensing into
3.2.4.1 DI
The DI fuel system, as shown schematically in Figure 3-3, applied a constant pressure to the
fuel through the use of a free piston accumulator and a compressed nitrogen bottle supplied
by British Oxygen Company (BOC), which was controlled using a high pressure gas
regulator manufactured by the company. The pressure (150 bar) supplied is more consistent
44
than that of a high pressure fuel pump. However, only 2 litres of fuel could be supplied to the
accumulator with each charge, necessitating the engine to be stopped at regular intervals in
The fuel required for the test was pumped using an electric out of tank pump from the main
fuel tank in the test cell, through a three-way valve (L) directly into the accumulator. The
three-way valve allowed the fuel to be pumped from the tank around the PFI system when it
45
was turned to the other position. A smaller auxiliary tank was used when a non-gasoline fuel
was tested, which allowed the gasoline to be constantly stored in the main fuel tank, rather
than having to be drained each time a non-gasoline fuel was used. The partially separate
systems also allowed contamination between the fuels to be eliminated to ensure that the test
data is reliable. The system operates via another three-way valve (M), so when the tank was
used, the valve was turned to its other position to allow fuel to be pumped from the auxiliary
fuel tank into the accumulator. Once the DI fuel system was charged, the pressurized fuel was
transported to the DI injector using a variety of rigid and flexible tubing, and stainless steel
compression fittings. The flow of fuel and nitrogen gas through the system was controlled
using various manually operated valves. When refilling the system and changing fuels, the
system was purged with nitrogen gas. If a non-gasoline fuel is used during the tests, the
auxiliary fuel tank and the system were cleaned using a small amount of the non-gasoline fuel
to ensure that there was no gasoline contamination in the non-gasoline fuel, when the fuel
tank was subsequently filled with it. Similarly, at the end of the day, the system was cleaned
with gasoline to avoid unnecessary damage to the injector and accumulator seals.
A low pressure (3 bar) PFI system, also shown schematically in Figure 3-4, was used for the
sole purpose of warming the engine prior to DI testing. As mentioned, the out of tank fuel
pump was operated to pump fuel from the gasoline fuel tank and the three-way valve (L) was
changed so that the fuel was pumped to the PFI injector. After the fuel exits the PFI injector,
it passed through a water cooling line in order to avoid excessive heating of the fuel and
excessive consequential vapour loss from the system, as shown in the figure.
46
3.2.5 DI Fuel Injector and Spark Plug Configuration
The combustion chamber of the engine has a flat top piston and a centrally mounted six-hole
direct injector along with a side-mounted spark plug located close to the injector. Figure 3-4
shows the relative positions of the injector along with the injector spray plume.
Figure 3-4 Relative positions of the direct injector along with the injector spray plume
The spark plug was located between fuel spray plumes 1 and 6, and it was inclined at an
angle of 18˚ to the cylinder axis. As shown in the figure, the injector, which was
manufactured by Bosch Ltd., had two symmetric groups of holes with three holes on each
Single-pulse injections were used throughout the experiments with an injection pulse width
ranging from between ~1.5 and ~3ms for the load range of 5.5-8.5 bar IMEP studied. The
injection event was triggered by a 12 volt DC electrical signal at the SOI of 280˚bTDC used
47
throughout the experiments. More information about the SOI and other electrical engine
3.2.6 EGR
A pipe with a valve was connected just after the exhaust plenum chamber to the intake
plenum chamber in order to implement and control hot EGR in the engine, as shown in
Figure 3-2. The Horiba MEXA-7100DEGR, discussed in detail in Section 3.5.2, had an in-
built function which allowed a sample line to be connected to it, enabling the system to
calculate the EGR addition (%) to the engine (accurate to ±0.02%). The sample line was
connected to the intake and the CO2 percentage was compared to the CO2 percentage in the
(%) (3-1)
The position of the EGR valve (±0.25% EGR accuracy) was then changed manually along
with the throttle angle and injection pulse width to achieve the desired load and EGR
addition. Although the intake manifold temperature was not controlled in the engine, the
effects of the increased intake manifold temperature with increased EGR addition have been
taken into account in the results. (Please note that whenever EGR is discussed in this thesis,
In order to achieve cooled EGR in the engine, the setup for the hot EGR was modified
slightly to incorporate a 1.1 metre long heat exchanger to cool the EGR gases before they
entered the engine, as shown in Figure 3-5. Air was blown over the heat exchanger to cool
the EGR gases as they passed through and the overall length of the EGR line was increased
48
slightly in order to incorporate the heat exchanger. Although some water vapour was found to
condense in the heat exchanger, it is believed that the vast majority did not condense. Thus,
the water vapour content of the EGR gases was not significantly changed. The same
procedure was followed as with the hot EGR to achieve the desired engine load and EGR
addition.
In order to simulate the effect of EGR after a TWC, a gas bottle supplied by BOC with a
mixture of 86% nitrogen and 14% CO2 was utilized. The mixture represents the composition
of gases that would be present after a TWC assuming that 100% of the NOx and CO from the
engine has been converted to N2 and CO2 respectively, and assuming that all of the water
vapour has been condensed and removed. It also assumes that all of the HCs and PM have
been successfully oxidized in the exhaust system. This removed the effect of varying TWC
exhaust gas conversion efficiencies which have been shown to vary the EGR ratio
(Hoffmeyer et al., 2009, Lewis et al., 2014). This is due to changes in the CO conversion
efficiency which affects the CO2 percentage of the exhaust gas, affecting calculation of the
49
Furthermore, it reduces the calorific value of the intake charge since energy cannot be
extracted by oxidizing the HCs and PM in the engine cylinder and it reduces the heat capacity
of the EGR gases because the water content of the EGR gases is removed. By using a
simulated TWC, the back-pressure effect of a real TWC was eliminated allowing the effects
of the post-TWC EGR gases on the engine combustion and emissions to be studied alone
without the back-pressure effects. Fuel consumption would be expected to increase if the
The gas bottle was controlled using a regulator manufactured by BOC and a flow rate
controller, after which a one-way valve manufactured by Swagelok was used to prevent any
possible backflow into the gas bottle. The gas then passed through a 5 metre heated line
maintained at temperatures between 75-200˚C, depending on the flow rate of gas through the
line, to simulate the high temperature of the EGR gases after a TWC. Finally, the heated line
was connected to the intake plenum on the engine to allow the EGR gases to mix sufficiently
with the intake gases before they enter the engine. This setup is shown schematically in
Figure 3-6.
50
In order to achieve the desired engine load and EGR addition, the same procedure was
followed as with the hot and cooled EGR. However, rather than the EGR valve being
adjusted along with the throttle and injection pulse width, the flow rate controller was
adjusted along with the throttle and injection pulse width and the temperature of the heated
line was adjusted accordingly. Higher temperatures were used for higher flow rates and lower
temperatures were used for lower flow rates, to achieve approximately the same temperature
of heated gas for all flow rate conditions tested, as measured by the intake plenum
temperature thermocouple. This was done to accurately reflect the similar EGTs that would
be experienced when testing at a single engine load for the different EGR types in this
investigation.
Although the EGT at a particular engine load will decrease with EGR addition, it is expected
that the differences would become negligible after the TWC in a real vehicle. This is because
sufficient heat transfer between the exhaust system and the atmosphere would be expected to
occur by the time the exhaust gases passed through the TWC, cancelling out any relatively
small EGT differences that would be otherwise observed at different EGR additions.
The same setup and procedure was used for the cooled EGR after TWC, except without the
51
Figure 3-7 Cooled EGR after TWC schematic
Baffle plates, as shown with drawings in Figures 3-9b and 3-9c, were installed on the engine
between the intake runner and cylinder head in order to achieve swirl and high tumble airflow
conditions, respectively, within the engine cylinder. The plates worked by blocking part of
the intake flow into the engine to generate the swirl and high tumble. The swirl and tumble
ratios were estimated using findings from other researchers, the results of which are presented
in Section 3.2.7.1. As well as these two plates, another baffle plate shown in Figure 3-8a was
made with the same profile as the intake runner, so that the intake geometry was the same for
the reference low tumble condition as it was for the swirl and high tumble conditions. The
tumble for this condition was generated by the vertical inclination of the intake runner alone.
Low tumble baffle plate Swirl baffle plate High tumble baffle plate
Figure 3-8 a) Low tumble baffle plate, b) swirl baffle plate, c) high tumble baffle plate
52
3.2.7.1 Swirl and Tumble Ratio Estimation
For this investigation, the swirl and tumble ratios of the single-cylinder engine with the baffle
plates fitted (Figure 3-8) were estimated based on data from previous investigations by other
researchers. Several attempts were made to measure the ratios using PIV in a single-cylinder
optical engine; however, they were unsuccessful due to technical difficulties which could not
be overcome within the time available. Therefore, it became necessary to estimate the ratios
Based on these findings of other researchers, the estimated swirl and tumble ratios of the
baffle plates, as shown in Figure 3-8, have been estimated as shown in Table 3-2. A summary
of the equipment and methods used by the other researchers to obtain their swirl/tumble ratios
Table 3-2 Estimated swirl and tumble ratios for the three intake airflow conditions
53
Table 3-3 Summary of other researchers findings for swirl and tumble ratios
Researcher Engine Cylinder Number Bore Intake Piston Engine Swirl/Tumble Swirl/Tumble Swirl Ratios Tumble
Type Head of Valves (mm) Runner Shape Speed Generation Measurement Ratios
Design Inclination (rpm) Method Method
Angle to
horizontal (˚)
AJ133 Engine Single- Pentroof 4 89 45 Flat Piston 1500 Baffle Plates - - -
cylinder
(Floch et al., Single- Pentroof 4 88 45 Flat Piston - Baffle Plates Steady-flow test rig 0 (no baffle 0.7 (no baffle
1995) cylinder plate), 1.5 (swirl plate), 1.8
generating (tumble
baffle plate) generating
baffle plate)
(Teraji et al., Single- Pentroof 4 93 30 Flat Piston 1500 Baffle Plates CFD (STAR-CD) 0 (no baffle 1.5 (no baffle
2009) cylinder plate), 1.5 (swirl plate), 1.5
generating (swirl
baffle plate), 0 generating
(tumble baffle plate),
generating 2.1 (tumble
baffle plate) generating
baffle plate)
(Fujimoto and 4-cylinder Pentroof 4 78 n/a (only swirl Flat Piston - Throttle plate in Steady-flow test rig 0.4 (slightly -
Tabata, 1993) (modified ratios measured, intake runner closed plate)-3.8
to operate and this only passage of one (fully-closed)
as single- affects tumble intake valve
cylinder) ratio)
(Li et al., 2001) 4-cylinder Pentroof 4 80 n/a (only swirl Flat Piston 1200 Blockage of intake PIV 2.30 -
(modified ratios measured, port into one intake (180˚bTDCcomb)
to operate and only this valve - 1.90
as single- affects tumble (60˚bTDCcomb)
cylinder) ratio)
54
Overall, it can be seen that the swirl/tumble ratios of other researchers have been observed
with equivalent or similar methods for generating the swirl and tumble intake airflows, as
compared to those used in this investigation, and they have been observed with similar 4-
valve pentroof combustion chamber designs. The results observed by other researchers are
similar to one another across similar engine designs (pentroof combustion chamber), with
different swirl/tumble generation methods and different intake runner inclinations to the
horizontal, accounting for the differences in swirl/tumble ratios observed. Therefore, while
not being 100% accurate, it is reasonable to expect that the swirl and tumble ratios in Table 3-
Furthermore, tumble ratios have been shown to vary depending on what equipment is used to
measure them (Kim et al., 2006). Therefore, it is believed that the estimation of the swirl and
tumble ratios obtained will not affect the interpretation of the findings of this investigation
more significantly than the use of different swirl/tumble measurement equipment. It is further
believed that the relative differences in swirl and tumble ratios between the baffle plate
designs have been accurately estimated, although the absolute magnitudes may differ slightly
from the true values. Thus, the interpretation of the findings of this investigation will not be
significantly affected.
55
3.2.8 Variable Compression Ratio
Although the single-cylinder engine had a geometric compression ratio of 11.5:1, the
compression ratio could be modified by adjusting the number and size of metal inserts which
are placed beneath the cylinder head. These acted to adjust the height of the cylinder head in
relation to the piston BDC, allowing the compression ratio to be changed. The maximum
compression ratio was however limited; beyond this limit the intake valves would have hit
the piston crown. The real dynamic compression ratio of the engine could be adjusted by
changing the valve timings; however, this was not pursued in this investigation. The
compression ratio of the engine is thus calculated using Equation 3-2 (Heywood, 1988).
where Vd is the swept volume which remains constant and Vc is the clearance volume which
is altered with the metal inserts. To achieve the desired compression ratio, the equation was
solved for Vc which could then be related to the new size of the metal inserts (h) using
Equation 3-3. This allowed the desired compression ratio to be achieved by inserting the
(3-3)
56
3.3 Control Systems
The spark timing and injection pulse width along with the start of injection and coil charge
duration were controlled using the engine timing control system (ETCS); a program written
using LABVIEW and operated using a NI card (model 6602). This program was initially
developed by the ICE group at the University of Oxford but subsequent modifications were
made by a previous research student in the engine group (Luszcz, 2009). It allowed the
In order to achieve a particular engine load, fuel was injected and adjusted using the ETCS
software and the airflow into the engine was controlled using a butterfly valve that was
adjusted from the engine control room. The fuel and airflow were modified until the desired
57
engine load was achieved. When EGR was used, the EGR valve was also adjusted along with
the injection pulse width and throttle position, as discussed, in order to achieve the desired
control injection, ignition and the variable valve timings. This was achieved using a timing
light which was attached to the ignition coil of the engine, enabling it to produce a
stroboscopic light signal when the ignition coil was activated by the ETCS system at a
specified ignition timing during low speed engine cranking. The light was shone onto the
flywheel which had CAD markings; the reading achieved when the light came on was
recorded and the encoder position was subsequently adjusted until the system was
synchronized.
positional offset of the intake and exhaust camshafts, relative to the position of the camshaft
previous research student in the engine group (Luszcz, 2009) which controls the valve timing
using a NI card (model 6202), was used. A screenshot of the program front-panel is shown in
Figure 3-10. A Hall effect type sensor was used to detect the relative position of the camshaft
and this position was controlled using hydraulic solenoids mounted in the cylinder head,
which adjusted the oil pressure and thus changed the relative camshaft positions. The
program constantly adjusted the oil pressure in order to maintain the cam timing within a
range of .
58
Figure 3-10 VVT LabView program front-panel (Luszcz, 2009)
Table 3-4 shows the geometry of the intake and exhaust camshafts and Figure 3-11 shows the
intake and exhaust valve lift profiles along with that of a typical in-cylinder pressure trace.
The camshaft itself had a high-lift profile typically used in SI engines and the valve timings
were chosen in order to minimise the residual gas fraction to increase the effect that EGR has
on the combustion and emissions. The engine was designed with an intake valve of a larger
inner seat diameter and higher maximum lift than the exhaust valve, in order to improve VE.
59
30
Intake Valve Lift Profile
Typical In-Cylinder
20 Pressure Profile
15
10
0
0 180 360 540 720
TDCIntake TDCCombustion
CAD
All of the signals sent from the control software to operate the engine as well as those
produced by the sensors on the engine were sent to the control tower which consisted of the
ETCS and VVT interface boxes which were used for the ETCS and VVT systems,
respectively. Figure 3-12 summarizes the arrangement and flow for the various signals.
60
Figure 3-12 Engine ETCS and VVT control systems; their arrangement and signal flow
Detailed instructions on how to set up the variable valve timing system can be found in
(Wang, 2014b) and (Daniel, 2012d). These instructions were followed in order to set up the
3.4 Instrumentation
Various ancillary equipment was fitted onto the research engine and balancing unit in order to
provide the measurements which were recorded and later analysed in this investigation.
These mainly consisted of pressure, temperature and engine speed data, each of which is
61
3.4.1 Engine Speed and Torque
The engine was coupled via a universal joint to a DC dynamometer in order to maintain a
constant speed of 1500 rpm ( rpm), regardless of the engine torque output. The speed was
chosen because it represents a speed used frequently in emission test cycles and it is
commonly used by researchers (Alger et al., 2009a, Heffel, 2003, Institution of Mechanical
Engineers, 2013). This along with the medium-high engine loads (5.5-8.5 bar IMEP) studied
one of the worst conditions for fuel consumption and emission formation. A torque meter on
the engine dynamometer was used to provide an instantaneous torque reading, which is useful
In-cylinder pressure measurements in the engine were made using a Kistler 6041A water
cooled piezoelectric pressure transducer along with a Kistler 5011B charge amplifier, which
was used to amplify the signal from the transducer. The transducer itself was fitted at the side
of the engine cylinder head between the intake and exhaust valves, to ensure that it had direct
contact with the gases in the combustion chamber for accurate in-cylinder pressure
measurement (Hountalas and Anestis, 1998). The electric charges induced by the
piezoelectric material were amplified by the charge amplifier and then outputted as a
modulated voltage. The pressure transducer was calibrated by Kistler using an oil weight
bench machine in the range of 1 MPa/1 V to 10 MPa/10 V, and this calibration was then
input into the charge amplifier as a tuning coefficient. Overall, the system was capable of
absolute pressure measurements in the range of 0.08-10 MPa. Once the signals were
62
In order to measure the intake and exhaust manifold pressures, EPT 3100 media isolated
thermocouples supplied by RS were used for the temperature measurements. The signals for
these temperature sensors were sent to a TCK-4 thermocouple amplifier before being
engine for fuel consumption calculations. The flow meter was designed to rotate one
revolution for a fixed volume of air, allowing the airflow rate to be quantified. The meter
itself was attached via a coupling to an encoder, with its speed recorded by the data
acquisition system. In order to relate the frequency of the encoder to the volumetric airflow
rate in the engine, a calibration was conducted at different throttle angles and thus different
engine volumetric airflow rates. The airflow in the engine was measured using an orifice
plate connected after the 100 litre intake damper and before the throttle, which created a unit
pressure difference for a unit flow across it. This pressure difference was measured using an
inclined u-tube manometer and the reading was used to calculate the actual flow rate in the
Cd is a dimensionless number representing the orifice plate discharge coefficient and is the
ratio of the orifice flow plate areas. Cd can be obtained by dividing the actual flow rate with
the theoretical flow rate. This actual flow rate was calculated at a known orifice flow rate
litres/second. The theoretical flow rate in this case was calculated using Equation 3-4 by
63
assuming that Cd is 1. The moisture pressure was not taken into account when calculating the
air density, limiting the accuracy of this method. Along with the u-tube monometer readings,
the frequency of the encoder was recorded, therefore allowing a correlation between the two
to be calculated. The calibration equation was then input into the low speed data acquisition
(LSDAQ) software, which is discussed further in Section 3.6.2, in order to calculate the
volumetric airflow rate at any given engine condition. The u-tube manometer and encoder
frequency readings were repeated three times at each throttle position in order to improve the
Figure 3-13 shows the calibration between the calculated and actual airflow rates. It is clear
that the calibration achieved is good. However, even with a R2 value of 0.9966, there is an
average percentage error of 1.6% between the calculated and actual airflow rates. This was
obtained by individually calculating the percentage error for each of the data points on Figure
3-13, and then taking the mean average of these. The error is largely due to the limited
resolution of the u-tube manometer device and the limited unit pressure difference produced
64
300
y = 1.0001x + 0.3181
250 R² = 0.9966
150
100
50
0
0 50 100 150 200 250 300
Actual Air Flow Rate (l/min)
The minimum and maximum VAF rates measured in this study were 200 and 300 l/min
respectively, which is within the calibration range shown in Figure 3-13. The maximum
airflow rate in the engine if the VE was 80% would be 339 l/min according to Equation 3-5.
(3-5)
Therefore, the maximum airflow rate used in the experiment of 300 l/min at an engine load of
8.5 bar is approximately correct, considering the engine can reach a maximum load of
LA4 Lambda meter and a Bosch heated wideband oxygen sensor. Fuel-specific curves were
used in the Lambda meter to interpret the dynamic (actual) AFR using the oxygen content in
the exhaust. Before a new fuel is tested in the engine, its hydrogen-to-carbon (H/C) and
oxygen-to-carbon (O/C) ratios were input into the system along with its stoichiometric
65
air/fuel ratio. This was then used by the Lambda meter to characterise the fuel curves to
calculate the dynamic AFR. As discussed, in order to achieve a specific engine load, the
throttle and injector pulse width were modified along with the EGR valve, if EGR was used.
Not only was the target load achieved using this procedure but the target Lambda value too.
Overall, this system was open loop meaning that the injection pulse width required regular
measured. The devices used to measure these emissions are described in the following sub-
sections.
plenum, and they were measured using two devices; firstly the differential mobility
spectrometer (DMS) 500 manufactured by Cambustion Ltd. and secondly the scanning
mobility particle sizer (SMPS) 3936, manufactured by TSi. Through experience, the DMS500
provided more consistent readings, resulting from the equipment being newer. However, due
to limited equipment availability, the SMPS 3936 was used in these experiments too.
In order to improve the consistency of the PN data, particles smaller than 10 nm which are
detectable by the PM measurement equipment used in this investigation were removed from
the range of particles studied. The particles have been characterized in this investigation by
their diameter in order to distinguish between ultra-fine particles and fine particles, since the
ultra-fine particles are the most significant contribution to PN and they are the most harmful
66
In this investigation, the nucleation mode particles have been classified as those within the
10-30 nm range and the accumulation mode particles have been classified as those within the
30-500 nm range for data from the DMS 500. For the results from the SMPS 3936, the
accumulation mode particle range was 30-300 nm, due to the lower maximum particle size it
can measure. This is based on the PM data recorded which can be seen to form two distinct
peaks within the stated ranges. (Zhang et al., 2014b) also classified nucleation mode particles
engine investigation using gasoline fuel along with butanol and ethanol splash blends.
It must be noted that the separation between the two modes is ill-defined (Kittleson, 1998)
and that in this investigation; the PM accumulation mode was shifted towards a smaller size
for the data collected using the SMPS 3936, because of the use of the thermodenuder, as will
be explained in a later section. It must also be noted that with the DMS 500 when no
thermodenuder was used and with the SMPS 3936 despite the use of the thermodenuder,
some nucleation mode particles still remained in the accumulation mode peak. This is
because they were adsorbed onto the accumulation mode particles. Also, despite the use of
the thermodenuder with the SMPS 3936, a nucleation mode particle peak was still observed.
The DMS 500 classified particles according to their size by separating them in a parallel
67
Figure 3-14 Cambustion DMS 500 analyser operating principle
Firstly, the particulates were pumped from the exhaust into the two stage dilution system
built into the instrument, as shown in Figure 3-15. The sample line from the exhaust to the
dilution system was maintained at 150˚C in order to avoid condensation of the particulates.
The particulates were then diluted in the primary dilution stage at a fixed dilution ratio of 5:1
for gasoline engine measurement, in which the mass flow rate is controlled by HEPA filtered
compressed air. A rotating disc was then used to perform the secondary high ratio dilution
with a dilution ratio of 16:1 used for this investigation; the system allowed a maximum
secondary dilution ratio of 500:1. This resulted in an overall dilution ratio of 45.2:1. The
choice of dilution ratios ensured that a good signal to noise ratio was maintained while
2011). Samples were collected at a rate of 1 Hz for approximately 2 minutes for each reading
68
Figure 3-15 Schematic of the DMS 500 two-stage dilution system
Once the particles were diluted, they were passed through a unipolar corona charger. This
charged the particles with a positive electric charge which was proportional to the particle
size. The filtered sheath flow then pumped the charged particles into the particle classifier.
The high voltage electrode as shown in Figure 3-14 then attracted the positively charged
particles towards the electrometer detector. The distance travelled by the particles along the
classifier was proportional to their size and inversely proportional to their mass. The classifier
which consisted of 22 electrometer detectors measured the charge quantity and distance
travelled by the particles, thus allowing the particle size, mass and number data to be
69
3.5.1.2 SMPS 3936
The SMPS 3936 system was comprised of three different units. Firstly, a TSi rotating disk
thermodiluter (model 379020A) maintained at 150˚C was used to dilute the exhaust sample
before it was pumped through a Topas TDD 590 thermodenuder at a temperature of 400˚C, to
remove most of the volatile nucleation mode particles in order to make it easier to distinguish
the nucleation and accumulation mode particles from one another. PM emissions from DISI
engines are mostly composed of volatile particles; they are fundamentally different from
those produced by diesel engines. Volatilities not only exist in the nucleation mode, but they
are also adsorbed onto accumulation mode particles. This makes it difficult to measure the
PM size distribution accurately. Therefore, the use of the thermodenuder device enabled the
Secondly, the particles were processed in the electrostatic classifier (model 3080) which had
a differential mobility analyser (model 3081) attached to it, in order to separate the particles
according to size. Thirdly, the particles were counted in the condensing particle counter
(model 3775). The settings inputted into the SMPS system for this study are summarized in
Table 3-5. The overall dilution ratio used in this study for the SMPS is 45.72:1, based on the
70
Table 3-5 SMPS 3936 system settings
In addition to making it easier to distinguish the nucleation and accumulation mode particles
from one another, the use of the thermodenuder also resulted in a significant PN reduction
and the PM distributions being shifted towards a smaller size. As discussed, both of these
were clearer at higher engine loads where there was more soot formation. This is due to the
non-adsorbed nucleation mode particles and those nucleation mode particles adsorbed onto
accumulation mode particles, respectively for the nucleation and accumulation mode
particles, being partially removed by the thermodenuder. This was a compromise that the
author and the previous research student (Wang, 2014b) made in order to improve the ability
to distinguish between the nucleation and accumulation modes, as mentioned. Overall, what
is most important is the relative change between different engine conditions rather than
absolute numbers, since these can vary significantly between different engines (Wang,
2014b).
to the PM sampling point, downstream of the exhaust plenum, before being pumped through
a pre-filter and heated line, both of which were maintained at 190˚C to avoid condensation of
the emissions. They were then subsequently analysed by the Horiba MEXA-7100DEGR. The
71
Table 3-6 Horiba MEXA-7100DEGR specification
Hot-wet FID
NDIR (dry)
NDIR (dry)
Accuracy 1%
Nitrogen oxides in the Horiba system were measured using a dry chemiluminescence detector
(CLD) and total hydrocarbons, HC, were analysed using a hot-wet flame ionization detector
(FID) (Robert Bosch GmbH, 2006). Prior to the NO x being analysed, the NO2 was converted
into NO to enable it to be measured in the CLD. Once in the CLD, ozone was used to
produce a reaction with the NO, and as a result a photon was emitted. This produced a current
which was converted to a voltage output which was a function of the NO and thus NO x
concentration in the sample. For HC analysis, the sample gas was pumped into a nozzle
charged with a high voltage, after which it was analysed using a hydrogen flame. In the high
temperature hydrogen flame, some of the hydrocarbon molecules were ionized, resulting in a
flow of current between the positively charged nozzle and a collector. The current was then
converted into a voltage output which was a function of the HCs in the sample. The voltage
outputs created by the respected CLD and FID detectors were used to calculate the relative
72
Oxides of carbon (CO and CO2) were measured using a non-dispersive infra-red (NDIR)
detector. The device measured the change in light intensity as an infra-red beam was passed
through the sample gas chamber. Oxides of carbon absorb infra-red light so the change in
light intensity could be used to calculate the relative concentrations of CO and CO 2 in the
sample gas.
Prior to each test, the Horiba was heated to its pre-set temperatures and the device was
calibrated using zero calibration gases. The test points were typically recorded one after
another during which the system continuously measured the gaseous emissions. When the
engine was stopped, the system was purged with compressed air to avoid build-up of
the low speed, time resolved engine data was also recorded using a NI card (model 6602).
In order to acquire detailed in-cylinder pressure data with a good resolution, the in-cylinder
pressure data was recorded at high speed with a resolution of 0.5 CAD for 300 consecutive
cycles, using the high speed data acquisition (HSDAQ) program written in LabView by a
previous research student in the engine group (Luszcz, 2009). The screenshot of the front
panel of the program is shown in Figure 3-16. With an engine speed of 1500 rpm, this
equates to data being collected every 80 ms (12.5 Hz). In order to ensure the correct CAD
was recorded for the in-cylinder pressure data, the intake pressure was also recorded and the
two were “pegged” to each other, as described in Section 3.6.4. An average was taken for
73
these data over 300 engine cycles to reduce the effect of small fluctuations in the
measurements on the data. The stability of the VVT system was measured by recording the
intake and exhaust cam position data from their respected sensors.
(intake, exhaust, cylinder wall, coolant, oil), ambient conditions (pressure, temperature), VAF
rate, throttle position, Lambda and the gaseous emissions data (NOx, HC), the data was
collected on a time resolved basis rather than a crank resolved basis. As with the high speed
data, an average was taken for these data, with an average taken over 200 samples. The data
was acquired using the LSDAQ program, again written in LabView by a previous research
student (Luszcz, 2009). The screenshot of the front-panel of the program is shown in Figure
3-17.
74
Figure 3-17 LSDAQ LabView program front-panel (Luszcz, 2009)
by previous research students at the University was used to analyse the data from the high
and low speed data acquisition systems. The method described by Stone (Stone, 1999) was
used to calculate the net IMEP from each engine cycle, and this was averaged across 300
consecutive cycles.
As mentioned, the intake pressure data was collected along with the in-cylinder pressure data
in order to ensure the correct CAD was recorded for the in-cylinder pressure data. The two
were “pegged” to one another at the same CAD, (BDCintake), where the two pressures were
75
equal, to ensure that the correct CAD was recorded for the in-cylinder pressure data. Overall,
this helped to overcome the effects of drift which would otherwise have made the
2015). A smoothing function was programmed into the MatLab script in order to reduce the
Therefore, a Ricardo WAVE model was created to simulate the in-cylinder conditions
recorded in the experiments, in order to calculate the average in-cylinder temperatures. The
model was then calibrated and verified using five cases for each, to ensure the model was
accurate for the engine conditions tested in this investigation. This is described in the
following sections.
Figure 3-18 shows the Ricardo WAVE model used in this investigation.
Ricardo WAVE is a one-dimensional engine and gas dynamics simulation tool created by
Ricardo PLC, and it is widely used in engine research and development by Universities and
the automotive industry. In order to simplify the analysis, the software uses the ideal gas law
combined with a prediction of trapped residuals and fuel vaporization behaviour, in order to
76
estimate the average in-cylinder gas temperature. This is much simpler than detailed chemical
Therefore, while the average in-cylinder temperature data in this investigation does not
represent actual measurements, it does provide some insight into the global average gas
temperatures. The fuel properties of indolene were used to simulate the combustion of
gasoline in order to further simplify the analysis. When simulating the combustion of the fuel
blends used in this investigation, the known properties were inputted but some unknown
combustion sub-model based on the recorded MFB profile was used along with a SI Wiebe
combustion sub-model, which required the input of MFB50 and MFB10-90, in order to
2. Engine geometry such as the bore, stroke, valve diameters and cam profiles were
measured
3. The information was then used to construct the Ricardo WAVE model geometry
4. The model self-check function was used to ensure correct matching of inlet and outlet
After the geometry had been produced, the other parameters were input into the software
after which it was calibrated and verified, as described in the following sections.
77
3.6.5.2 Ricardo WAVE Combustion Model Calibration
Five test points were used to calibrate the Ricardo WAVE model and the procedure shown in
The valve timings used for the model calibration were IVO=16˚bTDC and EVC=36˚aTDC,
and the IMEPs for the give cases are shown in Table 3-7.
78
Table 3-7 Test cases for Ricardo WAVE model calibration
IMEP (bar)
Case 1 3.63
Case 2 4.86
Case 3 6.19
Case 4 7.43
Case 5 8.68
The actual and calculated in-cylinder pressure data is shown in Figures 3-20a, 3-20b, 3-20c,
3-20d and 3-20e for the five respected cases. Overall, there is a good agreement between the
Fig. 3-21a: In-cylinder Pressure (case 1) Fig. 3-21b: In-cylinder Pressure (case 2)
40 40
In-cylinder Pressure (bar)
In-cylinder Pressure (bar)
35 35 Actual
Actual
30 30
Calculated Calculated
25 25
20 20
15 15
10 10
5 5
0 0
-50 0 CAD 50 100 -50 0 CAD 50 100
Fig. 3-21c: In-cylinder Pressure (case 3) Fig. 3-21d: In-cylinder Pressure (case 4)
40 40
In-cylinder Pressure (bar)
In-cylinder Pressure (bar)
35 35
Actual Actual
30 30
Calculated Calculated
25 25
20 20
15 15
10 10
5 5
0 0
-50 0 CAD 50 100 -50 0 CAD 50 100
79
Fig. 3-21e: In-cylinder Pressure (case 5)
40
30 Calculated
25
20
15
10
0
-50 0 CAD 50 100
Figure 3-20 In-cylinder pressure versus CAD data comparisons for Ricardo WAVE model
calibration: a) case 1, b) case 2, c) case 3, d) case 4 and e) case 5
The actual and calculated IMEP, Pmax, indicated efficiency and VE for the five cases are
1 5
0 0
Case 1 Case 2 Case 3 Case 4 Case 5 Case 1 Case 2 Case 3 Case 4 Case 5
80
Fig. 3-22c: indicated efficiency Fig. 3-22d: VE
0.50 0.7
Actual Actual
0.45
0.6
Calculated Calculated
0.40
0.35 0.5
0.30
0.4
0.25
0.3
0.20
0.15 0.2
0.10
0.1
0.05
0.00 0.0
Case 1 Case 2 Case 3 Case 4 Case 5 Case 1 Case 2 Case 3 Case 4 Case 5
Figure 3-21 Engine data comparison for Ricardo WAVE model calibration: a) IMEP (bar),
b) Pmax (bar), c) indicated efficiency and d) VE
The percentage difference for these parameters is shown respectively in Figures 3-22a, 3-22b,
3-22c and 3-22d for IMEP, Pmax, indicated efficiency and VE. As with the in-cylinder
pressure data, there is good agreement, showing that the model has been calibrated well. The
confidence intervals have been calculated and analyzed for the parameters above; these show
that the random error in the data was very low (less than 0.1% in most cases). Therefore,
random error is not thought to have significantly contributed to the differences observed
between the actual and calculated data, and because it is so small, it has not been displayed
5 5
4 4
3 3
2 2
1 1
0 0
Case 1 Case 2 Case 3 Case 4 Case 5 Case 1 Case 2 Case 3 Case 4 Case 5
81
Fig. 3-23c: indicated efficiency (% difference) Fig. 3-23d: VE (% difference)
6 6
5 5
4 4
3 3
2 2
1 1
0 0
Case 1 Case 2 Case 3 Case 4 Case 5 Case 1 Case 2 Case 3 Case 4 Case 5
Figure 3-22 Engine data percentage difference comparison for Ricardo WAVE model
calibration: a) IMEP, b) Pmax, c) indicated efficiency and d) VE
Overall, the VEs have all been calibrated to within 5% of each other at all of the engine loads,
and the other parameters correlate well with each other. Therefore, the calibration steps
Five further test points were used to verify the Ricardo WAVE model. The valve timings
used were IVO=6˚aTDC and EVC=61˚aTDC, and the IMEPs for the cases are shown in
Table 3-7.
IMEP (bar)
Case 1 3.74
Case 2 5.16
Case 3 6.47
Case 4 7.87
Case 5 8.90
The actual and calculated in-cylinder pressure data is shown in Figures 3-23a, 3-23b, 3-23c,
3-23d and 3-23e for the five respected cases. Overall, there is a good agreement between the
82
Fig. 3-24a: In-cylinder Pressure (case 1) Fig. 3-24b: In-cylinder Pressure (case 2)
35 35
25 Calculated
Calculated 25
20 20
15 15
10 10
5 5
0 0
-50 0 CAD 50 100 -50 0 CAD 50 100
Fig. 3-24c: In-cylinder Pressure (case 3) Fig. 3-24d: In-cylinder Pressure (case 4)
35 35
In-cylinder Pressure (bar)
In-cylinder Pressure (bar)
30 30
Actual Actual
25 25 Calculated
Calculated
20 20
15 15
10 10
5 5
0 0
-50 0 CAD 50 100 -50 0 CAD 50 100
83
Fig. 3-24e: In-cylinder Pressure (case 5)
35
25 Calculated
20
15
10
0
-50 0 CAD 50 100
Figure 3-23 In-cylinder pressure versus CAD data comparisons for Ricardo WAVE model
verification: a) case 1, b) case 2, c) case 3, d) case 4 and e) case 5
The actual and calculated IMEP, Pmax, indicated efficiency and VE for the five cases are
2 10
1 5
0 0
Case 1 Case 2 Case 3 Case 4 Case 5 Case 1 Case 2 Case 3 Case 4 Case 5
84
Fig. 3-25c: indicated efficiency Fig. 3-25d: VE
0.5 0.7
Actual Calculated Actual
0.4
0.6
Calculated
0.4
0.5
0.3
0.3 0.4
0.2 0.3
0.2
0.2
0.1
0.1
0.1
0.0 0.0
Case 1 Case 2 Case 3 Case 4 Case 5 Case 1 Case 2 Case 3 Case 4 Case 5
Figure 3-24 Engine data comparison for Ricardo WAVE model verification: a) IMEP
(bar), b) Pmax (bar), c) indicated efficiency and d) VE
The percentage difference for these parameters is shown respectively in Figures 3-25a, 3-25b,
3-25c and 3-25d for IMEP, Pmax, indicated efficiency and VE. As with the in-cylinder
pressure data, there is good agreement, showing that the model has been sufficiently verified.
However, there is an exception with cases 3 and 5 having a percentage difference larger than
5% for between the actual and calculated indicated efficiencies. Despite this, the difference
was still below 10% and the other parameters had a percentage difference less than 5%.
Again, the confidence intervals have been calculated and analyzed for the parameters above.
As with the calibration results, random error was found to be less than 0.1% in most cases.
Therefore, random error is not thought to have significantly contributed to the differences
observed between the actual and calculated data, and because it is so small, it has not been
85
Fig. 3-26a: IMEP (% difference) Fig. 3-26b: Pmax (% difference)
8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
Case 1 Case 2 Case 3 Case 4 Case 5 Case 1 Case 2 Case 3 Case 4 Case 5
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
Case 1 Case 2 Case 3 Case 4 Case 5 Case 1 Case 2 Case 3 Case 4 Case 5
Figure 3-25 Engine data percentage difference comparison for Ricardo WAVE model
verification: a) IMEP, b) Pmax, c) indicated efficiency and d) VE
Overall, the VEs have all been verified to within 5% of each other at all of the engine loads,
and the other parameters correlate well with each other, therefore, the model has been
sufficiently verified for the load conditions (5.5-8.5 bar IMEP) used in this investigation. This
will enable it to be used to accurately calculate the average in-cylinder temperatures for the
recorded measurements in this investigation. Furthermore, the ability to change the intake air
temperature has been utilized in the study of EGR, particularly with hot and cooled EGR, in
order to incorporate the effect of the intake air temperature on the average in-cylinder
temperatures. Despite the model being accurate, the author acknowledges that more work
86
could be done in order to improve the accuracy of the modelling, such as creating and
(3-6)
where (heat capacity ratio) is the ratio of specific heats which is approximated for different
parts of the engine cycle, (dV/dθ) is the rate of change of in-cylinder volume with respect to
CAD, (dP/dθ) is the rate of change of in-cylinder pressure with respect to CAD and (dQw/dθ)
is the cumulative net heat release rate. The heat release rate was then integrated in order to
(3-7)
The numerator of the equation is the integral of the net heat release rate between a specified
intake crank angle before the combustion process (i.e. CAD at MFB0) and the crank angle of
interest (e.g. CAD at MFB50), and the numerator of the equation is the integral of the net
heat release rate across the complete combustion window (i.e. between the CAD at MFB0
and MFB100). Key parameters were then obtained from the MFB curve; mainly the MFB50
along with MFB10-90, which are important in the interpretation of the combustion data.
In order to calculate the indicated efficiency (η) for this study, Equation 3-8 was used
87
(3-8)
where is the net work output from the engine, LHV is the lower heating value of the fuel
(assumes that all water from the combustion remains as vapour) and is the net mass
aforementioned VAF readings (converted into a mass flow rate, ) and the Lambda value
(3-9)
where Air/Fuel ratiostoichiometric is the stoichiometric air/fuel ratio and P i is the indicated power.
The intake temperature was also used to calculate the air density in order to improve the
accuracy of the VAF rate reading (converted into ), before it was used to calculate the
fuel consumption.
pressure trace, to calculate the knocking amplitude. The script read the pressure data and
applied a Butterworth 2nd order type filter to isolate the frequency range of 4-12 kHz, which
ensured that the first and second harmonic knocking frequencies from the engine remained
after the low and high frequency engine generated signal noise had been removed. It then
calculated the knocking amplitude from the amplitude of the filtered pressure trace. This
provided the knocking amplitudes from the pressure data, which allowed the KLMBT spark
timing to be quantified for each engine condition. The KLMBT was defined as the most
88
advanced spark timing in which 97% or less of the knock amplitudes were below 2 bar. The
maximum acceptable knock amplitude of 2 bar was chosen based on the work of (Mittal et
al., 2007). If the maximum brake torque (MBT) was reached before the KLMBT spark
timing, then this spark timing was defined as the KLMBT; an asterisk has been added to
denote that the spark timing was not knock limited. Using this method of finding the KLMBT
spark timing resulted in a variation margin of up to 10% for lower load (5.5 bar IMEP) and
5% for higher load (7.0 and 8.5 bar IMEP) conditions when the spark-timing was adjusted at
each testing point for the engine tests in Chapters 4 and 5 of this thesis. It is believed that this
should not affect interpretation of the data because the difference made by the margin in load
can be neglected.
All of the results in this investigation are reported at the KLMBT spark timings, or at the
MBT spark timing if the knock limit was not reached, because these are the optimum engine
operating points in order to achieve best fuel consumption and engine power (Zhu et al.,
2007). The engine of a vehicle will always be calibrated at the KLMBT/MBT (or at a set
spark retard from the KLMBT/MBT) for steady state warm running, if emissions are not a
concern, because customers typically purchase a vehicle based on its rated fuel consumption
and output power. Therefore, the author considered them to be the most appropriate operating
points at which to record the data, in order to maximize the impact of this research.
The HSDAQ LabView program discussed in Section 3.6.1 was later modified by the author,
as shown in Figure 3-26, to implement on-line filtering of the pressure trace, in order to
quickly identify the KLMBT spark timing before the engine data was recorded. This enabled
the tests in Chapter 6 of this thesis to be completed in a significantly shorter time than the
method described in the previous paragraph. It also enabled the discussed load variation to be
eliminated.
89
Figure 3-26 Modified HSDAQ LabView program displaying the on-line knock amplitudes
(3-10)
where ρemission represents the density of the gaseous emission, Emissionppm is the reading of
the emission from the Horiba system which was recorded by the data acquisition system,
is the volumetric flow rate of the exhaust stream and P i is the indicated power.
90
The molar mass of the exhaust stream was calculated based on the fraction of each
component (NOx, HC, CO, CO2, N2) and it was subsequently used to calculate the density of
(3-11)
where MEmission represents the molar mass of the emission (e.g. NO), P and T are the exhaust
stream pressure and temperature, respectively, and R is the gas constant. The calculation was
conducted separately for the NOx and HC emissions to provide the data presented in the
thesis.
For the data presented in this thesis, the averaged data from the 3 readings for each
measurement was plotted along with the 95% confidence intervals, where appropriate, in
order to enable the significant effects on the data to be identified. The confidence intervals
(3-12)
These confidence intervals only address the random errors that occur in the measurements;
system errors are not included. However, these are discussed further in the following section.
91
3.8 Fuel Properties
The properties of the fuels used in this investigation are shown in Table 3-9. Both gasoline
and ethanol were supplied by Shell Global Solutions, UK. The 1-butanol and heptane were
Heat of Vaporization ∆vapH (25˚C) (KJ/mol) ~37.34 52.05 42.36 34.6 34.6
Overall, the properties of the gasoline used represent the majority of gasoline fuel sold in
Europe and America, making it suitable for this investigation. Although a commercial
gasoline fuel was used for part of this investigation (first part of Chapter 4), it matched the
RON and ethanol content of the ULG95 fuel used in the remainder of the investigation, so it
is assumed that the fuel specification did not vary significantly from the one shown in the
table. Oxygenate content is thought to have varied, however, based on the results observed
and this is discussed further in a later section. The gravimetric content (GC) analysis of the
92
Table 3-10 Gasoline GC Analysis
Component % volume
Paraffin 43.9
Olefin 11.7
Naphthene 6.2
Naphthene, 1.6
Polynaphthene <0.1
Aromatics 26.9
to ensure that the data recorded from the engine was approximately the same as a previously
recorded accurate baseline. If there was a significant variation between the two then the cause
was identified and resolved before the baseline test was repeated. The procedure was repeated
until the engine results on a particular day conformed to the previously determined baseline
conditions. Before recording data, the engine cylinder block was warmed to 95˚C as
behaviour. Once the engine was recording reliable data, the test point was conducted and it
was repeated several times until 3 consistent sets of data were recorded, in order to ensure
that the data recorded was an accurate representation of the engine combustion and emissions
and to reduce the impact of day-to-day fluctuations on the results. This was then repeated to
collect all of the test points, which were required on a particular day.
93
3.10 Summary
Overall, this chapter has described the experimental test facilities used in this work. All of the
tests were conducted in a 4-stroke spray-guided DISI single-cylinder thermal engine. The
engine is typical of modern, while not state of the art, gasoline DISI engines. Finally, the data
Although the author benefitted from the engine setup created and developed by the various
previous research students at the University, several important improvements were made in
order to further develop and improve the setup. Firstly, the connection of the gasoline fuel
line directly to the fuel tank along with the incorporation of a three-way valve in the PFI fuel
line, in order to fill the DI accumulator directly from the gasoline fuel tank, improved safety
in the engine test cell and made fuel refilling easier. Secondly, the connection of the fuel
tank, previously used to fill up the DI system, to the main gasoline fuel tank, allowed residue
fuel from the DI system to be easily pumped out of the tank. This greatly minimized fuel
vapours from entering the atmosphere when the DI system was subsequently purged into the
tank. Thirdly, the creation of separate fuel tank for non-gasoline fuels made it easier to
prevent contamination of the gasoline and non-gasoline fuels. Finally, interference between
signal lines for the VVT system and water temperature control system were eliminated.
Previously when the mains water control valve was triggered open to cool the engine, the
VVT control system was sometimes affected, causing the valve timings to reset, meaning
engine rebuilds, VAF meter calibrations and ongoing work to maintain the engine test
facility.
94
Chapter 4
The Effect of EGR and its Type on Engine Combustion and Emissions
The aim of this chapter is to provide details of the combustion and emissions characteristics
of a DISI engine operated with EGR addition at different engine loads of 5.5, 7.0 and 8.5 bar
IMEP, and with different EGR types of hot EGR, cooled EGR, hot EGR after TWC and
cooled EGR after TWC at a single engine load of 7.0 bar IMEP.
4.1 Introduction
The main combustion parameters investigated in this chapter are the KLMBT spark timing,
in-cylinder pressure, calculated average in-cylinder temperature, MFB, fuel consumption and
EGT. PM, NOx and HC are the main emission parameters investigated.
There is an ever growing demand for reduction of NOx emissions from engines and for
are at the forefront of public perception. One way to achieve these demands is to use external
EGR in engines to suppress the temperature rise in the combustion chamber, reducing NO x
emissions and allowing the KLMBT spark timing to be advanced, as well as enabling the
throttle to be more opened, providing significant fuel economy benefits. Furthermore the
Euro VI emissions regulations which limit PN for the first time have increased interest in the
effect of EGR addition on engine particulates, with the potential of EGR coming under
question because it has been shown to have a negative effect on particulates (Gill et al., 2011,
Ma et al., 2014).
Overall, the author recognizes that the Euro VI emissions regulations have led many to
question the potential of EGR, noting that it is generally considered to increase engine
95
particulates. Thus, this is an area that has been researched further in this chapter. There is also
a lack of work regarding the effect of different types of EGR (hot EGR, cooled EGR, hot
EGR after TWC, and cooled EGR after TWC) on engine combustion and gaseous emissions,
on the same engine at the same load condition, thus, this was also investigated further in this
chapter.
conduct the engine tests, to collect the data presented and discussed in this chapter. Relative
air-fuel ratio λ was maintained at 1 during the experiments and a COV IMEP of 5% was not
exceeded. Valve timings were set at IVO=16˚bTDC, EVC=36˚aTDC. KLMBT spark timings
and a fixed geometric compression ratio of 11.5 were used. In order to achieve EGR in the
engine, and in order to achieve the desired engine load and EGR condition, the equipment
setup and test procedure outlined in Section 3.2.6.1 was followed for the hot EGR condition,
and those outlined in Sections 3.2.6.2, 3.2.6.3 and 3.2.6.4 were followed for the cooled EGR,
hot EGR after TWC and cooled EGR after TWC conditions, respectively.
The load range of 5.5-8.5 bar IMEP was chosen to study in the first part of this chapter
because it represents the medium-high load range of this engine where EGR should be of
most benefit to reduce engine knock, thus making the results more useful. In addition, the use
of DI in the engine is most relevant to this load range for knock suppression; at lower loads
the reduced in-cylinder pressure prior to ignition resulting from the charge cooling is not
necessary and it reduces indicated efficiency. Furthermore, in a real engine, a DI fuel pump
will consume more energy than a PFI fuel pump. Therefore, there must be a fuel consumption
benefit of using DI; i.e. charge cooling, in order to compensate for the additional energy the
DI pump would consume. The load of 7.0 bar IMEP was chosen to study in the second part of
96
this chapter because allowed a reasonable, while not a high EGR addition of 14% to be
achieved. A higher engine load would have meant that the throttle would have needed to be
more opened, thus reducing the maximum level of EGR that would have been achieved. One
load was chosen to study so that an in-depth understanding could be formed, which would not
have been possible if multiple loads were studied with the multiple EGR types.
The spark was swept for all engine loads and all EGR types in order to find the KLMBT
spark timing using the technique outlined in Section 3.6.9, or by using the torque calculated
from the engine data if the MBT was reached before the knock limit. A mixture of pump and
research grade gasoline was used in this research with the properties outlined in Section 3.8.
The pump grade gasoline was used in the experiments for the first part of this chapter and the
research grade gasoline was used in the experiments for the second part of this chapter. The
pump grade gasoline specification did not differ significantly from the research grade
gasoline specification; the RON and ethanol content were the same. Therefore, it is believed
that the use of both gasoline types did not significantly affect the interpretation of the results.
Oxygenate content is thought to have varied, however, based on the results observed and this
is discussed further in a later section. The DMS 500 was used to measure the PM emission
and the Horiba MEXA-7100DEGR was used to measure the gaseous emissions of NOx and
HC. PM was not measured in the second part of this chapter because of limited equipment
availability.
The test matrix provided in Table 4-1 was carried out in order to investigate the effect of hot
EGR at different engine loads on the engine combustion and emissions. The results and
97
Table 4-1 Experiment test matrix (EGR addition & engine load)
IMEP (bar)
5.5 1 2 3 4 5
7.0 6 7 8 9
8.5 10 11 12
The test matrix provided in Table 4-2 was carried out in order to investigate the effect of
different EGR types on the engine combustion and gaseous emissions at the single load of 7.0
bar IMEP.
Table 4-2 Experiment test matrix (EGR addition & EGR type)
EGR Type
Hot EGR 1 2 3
Cooled EGR 4 5
Hot EGR after TWC 6 7
Cooled EGR after TWC 8 9
The KLMBT spark timings for the tested engine conditions are presented in Table 4-3. EGR
enabled the spark to be advanced significantly, through the suppression of temperature rises
in the end-zone. It can be seen in Figures 4-2a, 4-2b and 4-2c that EGR addition led to
calculated average in-cylinder temperature reductions even with spark advances for the
engine loads tested. However, the spark advances did enable the combustion temperature
reducing effect of EGR to be mitigated, as can be most clearly seen in Figure 4-2b for the
engine load of 7.0 bar IMEP. It must be noted that knocking could have been avoided if the
compression ratio was fully optimized for the fuel used. However, this was not pursued in
98
order to investigate the effect of the parameter changes on engine knock limit and the
For the load of 5.5 bar IMEP, the spark timings were not knock limited due to the low
combustion pressure and temperature as shown in Figures 4-1a and 4-2a respectively,
allowing them to be advanced by 12 CAD to maintain their optimum MBT phase across the 0
to 8% EGR range. However, the laminar flame speed reducing effects of EGR (Rhodes and
Keck, 1985) resulted in the combustion phasing becoming retarded from the optimum with
further EGR additions of 12 and 13% (Figure 4-3a), despite the further 6 CAD spark advance
The spark timings at the other engine loads however were knock-limited so they could only
be advanced by 8 and 4 CAD respectively for the engine loads of 7.0 and 8.5 bar IMEP. EGR
was limited to 3% for the engine load of 8.5 bar IMEP because the throttle valve was more
open (~15˚ throttle angle) compared to the other engine loads where the throttle valve was
more closed (~7˚ at 7.0 bar IMEP, ~5˚ at 5.5 bar IMEP), which increased the intake manifold
pressure and subsequently reduced the pressure difference between the exhaust and the
intake. This meant that the EGR gas flow rate in the EGR line was reduced, which reduced
the EGR ratio compared to the other engine loads. The relatively small throttle angles
resulted from the oversized throttle; it was designed for a 3.0 litre engine, meaning that it did
not need to be opened fully in order to achieve the maximum engine load, as the case would
have been with a smaller throttle valve. Despite the limited EGR that was achieved at the
highest considered load, it was still important to study this engine load to quantify what
improvements could be made with the limited EGR available. This condition is used to
represent the high-end load of this engine. The highest load of ~9.5 bar IMEP which can be
99
achieved with this engine was not pursued because no EGR addition can be achieved at this
It is anticipated that the KLMBT spark timing advances would have been greater for the
engine loads of 7.0 and 8.5 bar IMEP if the intake manifold temperature increases resulting
from the EGR addition could have been eliminated. This is because the lower intake
temperature would have decreased the engine knock tendency. For 5.5 bar IMEP, the flame
would be expected to travel slower in the cooler fuel-air mixture resulting from the cooler
intake manifold, thus necessitating greater MBT spark timing advances to attempt to maintain
Table 4-3 KLMBT spark timings (˚bTDC) (* = not knock limited) (EGR addition &
engine load)
IMEP (bar)
5.5 27* 33* 39* 45* 45*
7.0 14 14 18 22
8.5 10 12 14
The in-cylinder pressures for the engine load of 5.5 bar IMEP are presented in Figure 4-1a.
They show that in-cylinder pressure reduced as the EGR addition increased. This is because
the EGR gases had a higher heat capacity than fresh air and the flame travelled at a slower
speed (Rhodes and Keck, 1985), reducing the temperature and subsequent pressure rises in
the combustion chamber, causing the overall in-cylinder pressures to reduce. In addition,
there was up to a 10% load variation at the lower engine loads when the spark-timing was
adjusted at each testing point, as discussed, accounting for some of the observed in-cylinder
100
pressure reduction as EGR was increased; IMEP reduced by 0.15 bar across the EGR range at
The in-cylinder pressures for the load of 7.0 bar IMEP in Figure 4-1b show a different
behaviour to those at 5.5 bar IMEP, with slight increases observed, despite similar ratios of
EGR being achieved. This occurred because the spark timings for the engine load of 5.5 bar
IMEP were not knock limited, however, at 7.0 bar IMEP they were. Therefore, the spark
timings at 5.5 bar IMEP could always be advanced to attempt to continue delivering their
optimum MFB50 phase of 8-10˚aTDC, which resulted in the optimum in-cylinder pressure
and temperature at the 0% EGR condition (optimum balance between compression and
expansion work produced on the piston). With the EGR addition, the combustion became
slower, and this could not be compensated for by advancing the MFB50. This meant that the
EGR addition resulted in the overall combustion pressures being reduced. The spark timing at
7.0 bar IMEP however, being knock limited, could always be advanced towards its optimum
MFB50 phasing with EGR addition, causing the in-cylinder pressure to be maintained and
even increased slightly. Despite this, it appears that the average in-cylinder temperature could
not be maintained causing it to reduce across the EGR range as shown in Figure 4-2b.
Similar in-cylinder pressure behaviour was observed at the engine load of 8.5 bar IMEP in
Figure 4-1c along with similar in-cylinder temperature decreases across the EGR range as
shown in Figure 4-2c. Therefore, it is again concluded, as with 7.0 bar IMEP, that although
the KLMBT spark timings advances (towards their optimum MFB50 phasing) were not
sufficient to maintain or increase the average in-cylinder temperature, they were sufficient to
increase the in-cylinder pressures. Thus, EGR addition appears to reduce the calculated
101
At 5.5 bar IMEP, the in-cylinder pressure increases would likely have been reduced, due to
the lower intake temperatures, if the intake manifold temperature increases could have been
eliminated. It would be anticipated that the in-cylinder pressure increases would have been
greater for the engine loads of 7.0 and 8.5 bar IMEP if the intake manifold temperature
increases resulting from the EGR addition could have been eliminated. This is because the
40
35
30
25
20
15
10
0
-50 0 CAD 50 100
0% EGR 4% EGR 8% EGR
12% EGR 13% EGR
40
35
30
25
20
15
10
0
-50 0 CAD 50 100
0% EGR 4% EGR 8% EGR 12% EGR
102
Fig. 4-1c: In-cylinder Pressure (8.5 bar)
45
35
30
25
20
15
10
0
-50 0 CAD 50 100
0% EGR 2% EGR 3% EGR
Figure 4-1 In-cylinder pressures versus CAD at KLMBT spark timings for a) 5.5 bar
IMEP, b) 7.0 bar IMEP and c) 8.5 bar IMEP
The calculated average in-cylinder temperatures decreased across the EGR range for the three
tested engine loads of 5.5, 7.0 and 8.5 bar IMEP, as shown in Figures 4-2a, 4-2b and 4-2c. As
explained, despite the in-cylinder pressures increasing across the EGR range for the engine
loads of 7.0 and 8.5 bar IMEP, the calculated average in-cylinder temperatures decreased, as
they did at 5.5 bar IMEP (where in-cylinder pressure decreased across the EGR range), due to
the temperature suppression effect of the EGR gases, resulting from their higher heat capacity
At 5.5 bar IMEP, it would be expected that the in-cylinder temperature decreases would have
been greater if the intake manifold temperature increases could have been eliminated, due to
the lower intake gas temperature. However, for the engine loads of 7.0 and 8.5 bar IMEP it is
believed that the increased intake manifold temperatures resulting from the EGR addition
were compensated for by the less optimized combustion phasing resulting from the
103
consequent reduced KLMBT spark timing advance, compared to the maximum spark
advance possible. This was a result of those higher intake manifold temperatures; lower
temperatures would have enabled a greater spark advance before significant levels of knock
occurred. The gas temperature during the intake stroke would have been reduced slightly, if
the intake manifold temperature increases could have been eliminated. However, it is thought
that the improved KLMBT spark timing would have quickly compensated for the decrease by
the time temperatures reached the beginning of the range reported (500˚C), due to the
combustion happening in a smaller volume; thus, the effect would have been minimal.
1900
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
0% EGR 4% EGR 8% EGR
12% EGR 13% EGR
104
Fig. 4-2b: In-cylinder Temperature (7.0 bar)
2100
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
0% EGR 4% EGR 8% EGR 12% EGR
1900
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
0% EGR 2% EGR 3% EGR
Figure 4-2 Calculated average in-cylinder temperatures versus CAD at KLMBT spark
timings for a) 5.5 bar IMEP, b) 7.0 bar IMEP and c) 8.5 bar IMEP
105
4.3.1.3 MFB
From Figure 4-3a it can be seen that the MFB profile for the engine load of 5.5 bar IMEP
became smoother and elongated with EGR addition, as the laminar flame speed was reduced
by the EGR addition (Rhodes and Keck, 1985), increasing the combustion duration. This
decreased the rate at which the fuel was burned, resulting in the observed behaviour. This is
reflected in the calculated average in-cylinder temperature data discussed in the previous
section which showed that the overall combustion temperatures reduced with EGR addition,
not only at 5.5 bar IMEP, but at the other engine loads too. If the compression ratio was
optimised for the fuel used, the combustion phasing would have been optimised for each test
point. Therefore, it would be expected that the combustion of gasoline would have been
quicker at the knock limited spark timings due to higher in-cylinder turbulence at the time of
ignition.
The MFB for the load of 7.0 bar IMEP, as shown in Figure 4-3b, followed the same trend as
that for 5.5 bar IMEP between the EGR ratios of 0 and 4%; the profile became smoother and
elongated, as the combustion duration was increased. However, the MFB profile for the EGR
ratio of 8% became more like that observed at 0% EGR, and the combustion duration became
slightly longer. Despite this, the differences between the MFB data at the EGR additions of 4
and 8% are likely due to experimental uncertainties. The KLMBT spark timing was advanced
by 4 CAD between the EGR ratios of 8 and 12% resulting in the combustion beginning
earlier, allowing a more advanced MFB profile to be achieved as compared to that observed
at 0 and 8% EGR addition. However, the profile became more elongated because the slower
flame speed reduced the rate at which the fuel was burned. These results suggest that the
KLMBT spark timing advance achieved with EGR addition at 7.0 bar IMEP was more able to
overcome the reduced laminar flame speed caused by the EGR addition, but at 5.5 bar IMEP,
the spark advance was less able to overcome it. Overall, the gradients were much more
106
consistent for 7.0 bar IMEP (Figure 4-3b) as compared to 5.5 bar IMEP (Figure 4-3a)
because of this.
The MFB data in Figure 4-3c for the load of 8.5 bar IMEP showed a similar trend to the data
at 7.0 bar IMEP after the initial EGR addition and spark advance. However, as the spark was
advanced further with more EGR addition, the MFB profile became significantly advanced
suggesting that the increased spark advance at 3% EGR was better able to overcome the
combustion slowing effects of the EGR gases than was the case with the spark advance
achieved at 2% EGR.
Taking the MFB data for the three engine loads into consideration, the KLMBT spark
advances achieved with EGR addition could overcome the resulting slower flame speeds in
some cases, but in other cases, the advances were not sufficient, causing the MFB profiles to
become smoother and more elongated. This may be further related to the observed
improvements in fuel economy and the reduction in EGT, along with the engine emissions;
all of which will be discussed further in the following sections. Overall, these relationships
For 5.5 bar IMEP, the MFB profile would be expected to become more elongated if the
intake manifold temperature increases resulting from the EGR addition could have been
eliminated, due to the slower flame speed in the cooler fuel-air charge. This would be
expected to occur at the other engine loads too, but much less significantly, because the
temperature decreases with the cooler intake gases would be expected to be quickly cancelled
out by the effect of the greater KLMBT spark timing advances, as discussed. Furthermore,
the MFB profile for 5.5 bar IMEP between the EGR ratios of 0 and 8% would not become
more advanced or retarded, because the optimum combustion phasing would be expected to
107
be maintained. However, it would still be expected to become retarded at the EGR additions
of 12 and 13%. It would be anticipated that the MFB profile advances would have been
greater for the engine loads of 7.0 and 8.5 bar IMEP. This is because the KLMBT spark
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
0% EGR 4% EGR 8% EGR
12% EGR 13% EGR
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
0% EGR 4% EGR 8% EGR 12% EGR
108
Fig. 4-3c: MFB (8.5 bar)
1.0
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
0% EGR 2% EGR 3% EGR
Figure 4-3 MFB versus CAD at KLMBT spark timings for a) 5.5 bar IMEP, b) 7.0 bar
IMEP and c) 8.5 bar IMEP
Gravimetric ISFCnet at the KLMBT spark timings for the three engine load conditions at the
different EGR ratios tested is shown in Figure 4-4. There is a clear trend of decreasing fuel
consumption as the EGR addition was increased because, as explained previously, the
temperature rises in the end-zone of the combustion chamber, inhibiting knock that would
otherwise have been produced. This allowed the MFB50 phasing of the engine to be
advanced towards its optimum phase for the engine loads of 7.0 and 8.5 bar IMEP, increasing
the engine efficiency and consequently reducing the fuel consumption (de O. Carvalho et al.,
2012). This occurred because more of the combustion process occurred at a lower in-cylinder
volume and thus, it generated higher combustion pressures. This improved the rate of
combustion and it increased the expansion of the combustion products into useful energy.
109
Also, because the gas from EGR takes a part of the intake air volume, the throttle valve had
to be opened more to provide the same amount of fresh air as in the non-EGR case, or in
other words, the same amount of oxygen for combustion. This reduced the pumping losses
across the EGR range and thus the fuel consumption at all engine loads; PMEP reduced from
0.45 to 0.38 bar for the load of 5.5 bar IMEP, 0.33 to 0.25 bar for the load of 7.0 bar IMEP,
and 0.24 to 0.22 bar for the load of 8.5 bar IMEP, across their respected EGR ranges. For the
loads of 5.5, 7.0, and 8.5 bar IMEP, the fuel consumption reduced by 2.2, 4.1 and 1.0%
respectively across the EGR range. Furthermore, the reduced combustion temperatures with
EGR addition improved thermal efficiency (Ratnak et al., 2015, Siokos et al., 2015),
The rate of fuel consumption decrease at 5.5 bar reduced at the higher EGR ratios because the
spark timings were not knock limited for this relatively low load. Thus, between 0 and 8%
EGR addition, the spark timing was at its optimum phase, and it became retarded slightly
from its optimum phase as the EGR addition was increased to 12 and 13%. This left little
Furthermore, because the MFB50 became later at 12 and 13% EGR addition (Figure 4-3a),
the rate of fuel consumption decrease was reduced at these conditions. It is proposed that the
increased rate of fuel consumption decrease observed at the engine load of 7.0 bar IMEP
when the EGR ratio was increased from 8 to 12% occurred because the in-cylinder pressure
increased significantly in comparison to that observed with the previous EGR increase from 4
to 8% (Figure 4-1b). This was possibly due to the earlier start of heat release (Figure 4-3b).
The in-cylinder pressure at 8.5 bar IMEP only increased slightly between the EGR ratios of 0
and 2%, despite the EGR addition and consequent spark advance, leading to the relatively
small improvement in fuel economy (Figure 4-4). However, the in-cylinder pressures
110
increased significantly as the EGR addition was increased further to 3%, leading to the
For the load of 5.5 bar IMEP, the fuel consumption improvements would have worsened if
the intake manifold temperature increases could have been eliminated, because the
combustion duration would have increased, as discussed, meaning that the conversion
efficiency of the pressure and heat from the combustion would have reduced. Fuel
consumption improvements for the loads of 7.0 and 8.5 bar IMEP would have been greater,
due to the greater KLMBT spark timing advances that would be expected. This effect would
be expected to overcome the slightly increased combustion duration that would also be
Despite the fuel consumption reduction observed and discussed, it must be noted that the
VAF had an average error of 1.6%; therefore, the reduction of 1.0% observed across the EGR
range at the engine load of 8.5 bar IMEP is limited, because it could have resulted from this
error. The reductions observed for 5.5 and 7.0 bar IMEP are however above 1.6%, thus, they
are not limited. The error mainly resulted from the limited resolution of the u-tube
manometer device used to calibrate the air flow meter. Despite the resolution being 1 mm, it
still remained a limiting factor in achieving better calibration between the VAF and u-tube
manometer reading, thus limiting the VAF accuracy. The error also resulted from the orifice
111
267
265
263
ISFCnet (g/kWh) 261
259
257
255
253
251
249
247
245
0 2 4 6 8 10 12 14
EGR (%)
5.5 bar 7.0 bar 8.5 bar
Figure 4-4 Gravimetric ISFCnet versus EGR ratio at KLMBT spark timings
4.3.1.5 EGT
The EGT decreased significantly across the EGR range for the three tested engine loads, as
shown in Figure 4-5. Overall, it decreased by 21.9, 32.7, and 7.6˚C for the loads of 5.5, 7.0
and 8.5 bar IMEP, respectively. As previously discussed, the calculated average in-cylinder
temperatures significantly reduced with EGR addition for the engine load of 5.5 and 8.5 bar
IMEP, and they reduced less significantly for the engine load of 7.0 bar IMEP, contributing
to some of the reductions in EGT across the EGR range. However, it is thought that the main
contribution to the reduced EGTs was the advanced combustion phasing enabled by the EGR
addition. It is proposed that this led to a more efficient conversion of the pressure and heat
from the combustion into piston work, causing the gases to cool more in the engine cylinder
before they were ejected into the exhaust. With the combustion starting earlier as the EGR
addition was increased, this effect became greater, causing the EGTs to reduce across the
EGR range. For turbocharged engines with EGR, this effect is very useful for protecting the
turbocharger turbine.
112
The temperature decrease is lower for 5.5 bar IMEP compared to 7.0 bar IMEP despite the
maximum level of EGR increasing. This is because the conversion efficiency of the pressure
and heat from the combustion into piston work could not be improved due to the combustion
already being at its optimum phasing. Furthermore, the optimum combustion phasing could
not be maintained at 12 and 13% EGR addition, as discussed, resulting in the EGTs not
decreasing further at these conditions. The EGT at 7.0 bar IMEP increased slightly by 6˚C at
the EGR ratio of 4%, whereas one would expect it to have decreased. Another slight increase
was observed between the EGR ratios of 0 and 2% at 8.5 bar IMEP. Despite these observed
increases, they are not considered to be significant; therefore, they will not be commented on
further.
For the load of 5.5 bar IMEP, the EGT decreases would have likely reduced if the intake
manifold temperature increase could have been eliminated, due to the aforementioned
reduced conversion efficiency of the pressure and heat from the combustion, due to an
expected increased combustion duration. The EGT decreases would likely have been greater
for the engine loads of 7.0 and 8.5 bar IMEP, due to an increased conversion efficiency of the
pressure and heat from the combustion, resulting from the greater KLMBT spark timing
advances that would be expected. This effect would be expected to overcome the slightly
increased combustion duration that would also be expected at these engine loads.
113
680
660
640
EGT (°C)
620
600
580
560
0 2 4 6 8 10 12 14
EGR (%)
5.5 bar 7.0 bar 8.5 bar
4.3.1.6 PM Emissions
The nucleation mode of the PM emissions decreased and the accumulation mode increased as
EGR ratio was increased from 0 to 12% for the engine load of 5.5 bar IMEP, as shown in
Figure 4-6a. The two modes could be distinguished from one another because they formed
two distinct peaks; the nucleation mode peak is comprised of the smaller particles on the left-
hand side of the plot (10-30 nm) (Figure 4-6a) and the accumulation mode peak is comprised
of the larger particles on the right-hand side of the plot (30-500 nm). According to (Zhang et
al., 2014b) the nucleation mode particles mainly result from droplets formed by hydrocarbon
condensation and the accumulation mode particles are mainly composed of carbonaceous
At the engine load of 5.5 bar IMEP, the increase in EGR suppressed the temperature rise in
the combustion chamber, reducing the oxidation rate of soot particles, leading to the
increased accumulation mode particles observed between 0 and 12% EGR addition. This
appears to have been more significant than the decreases in primary carbon particles formed
114
by the thermal pyrolysis and dehydrogenation reaction of fuel vapour/droplets, which would
have occurred with the reduced in-cylinder temperatures (Zhang et al., 2014b). It is proposed
that the observed reductions in nucleation mode particles occurred because they were
adsorbed onto the increasing amounts of soot accumulation particles. Thus, as the soot
accumulation mode particles increased across the EGR range, the numbers of nucleation
mode particles decreased. In addition to the reduced oxidation rate of soot particles across the
EGR range caused by the reduced combustion temperatures, it is proposed that the spark
timing advances reduced the time available for fuel-air mixing. This led to poorer mixture
preparation and thus more pockets of high local equivalence ratio which could not burn
completely contributing to the higher numbers of soot accumulation mode particles observed
across the EGR range. The EGT decreases across the EGR range also contributed to the
Increases in nucleation and accumulation mode particles were observed across the EGR range
for 7.0 bar IMEP as shown in Figure 4-6b. However, the increase in the nucleation mode was
not consistent, therefore, it will not be commented upon further. It is thought that the
increases. Again, this appears to have been more significant than the decreases in primary
carbon particle formation (Zhang et al., 2014b). In addition, due to the spark timing advances
achieved, it is proposed that the mixture preparation would have been poorer, contributing
further to the increased soot accumulation mode particles observed across the EGR range.
This is reflected in the MFB data (Figure 4-3b) which shows that the MFB50 was advanced
significantly at the EGR ratio of 12% when compared to the MFB at 0% EGR. Again, the
EGT decreases across the EGR range also contributed to the accumulation mode PM
increases observed.
115
The observed behaviour at 5.5 and 7.0 bar IMEP was not observed at 8.5 bar IMEP in Figure
4-6c, despite the calculated average combustion temperature decreasing (Figure 4-2c). At 8.5
bar IMEP, there was no significant change in the nucleation and accumulation mode particles
observed. This is due to the low EGR addition achieved at this engine load.
Overall, the contribution of nucleation mode particles became weaker as the engine loads
were increased. This agrees with the reduced HC emission observed as the engine load was
increased (Figure 4-8); the HC emission is mainly composed of the volatile particulate
fraction included in the nucleation mode (Costagliola et al., 2013). Thus, this explains the
strong relationship between the HC emission and nucleation mode particles observed in the
data. The accumulation mode particles increased with engine load due to the higher rate of
soot production and growth at higher temperatures in the combustion chamber (Zhang et al.,
2014b). In addition, as the load is increased, more fuel is injected which results in more
locally rich areas in the combustion chamber, further increasing soot formation (Wang,
2014b). (Wang et al., 2014c) observed the activation energies and temperatures required to
oxidise soot increased with engine load, further explaining the increased accumulation mode
particles as the engine load was increased. Furthermore, there is increased fuel impingement
at higher loads resulting from the longer injection pulse width and increased fuel mass
The PM behaviour at 5.5 and 7.0 bar IMEP is different from that observed by (Alger et al.,
2010) and (Hedge et al., 2011) where significant soot reductions across the EGR range was
reported, with soot levels being represented by the accumulation mode. It is likely however
that because the engines in their investigations were turbocharged, the EGR addition reduced
the need to enrich the fuel charge for the purpose of reducing turbocharger inlet temperatures,
consequently reducing their soot emission. Clearly there are complex behaviours regarding
116
the formation of PM emissions in engines which need to be investigated in detail on a
fundamental level.
It is proposed that the decrease in accumulation mode particles at 5.5 bar IMEP as the EGR
addition was increased to 13% occurred because the spark timing (Table 4-3) was not
advanced as compared to that at 12% EGR addition. This means that while the mixture
preparation was equivalent, the combustion duration was increased (Figure 4-3a), providing
more time for the particles to oxidize in the hot flames. Additionally, there is a decrease in
the accumulation mode particles for the load of 7.0 bar IMEP at 4% EGR addition (Figure 4-
6b). It is believed that the increased exhaust temperatures (Figure 4-5) increased the rate of
particles from a DISI engine can be effectively oxidised between 400-600˚C (Wang et al.,
2014c).
If the intake manifold temperature increases could have been eliminated, the accumulation
mode particles would have been increased more significantly at 5.5 bar IMEP. This is due to
the expected increased MBT spark timing advances further worsening mixture preparation,
and the more significant in-cylinder temperature decreases further reducing the oxidation rate
of the particulates. Nucleation mode particles would consequently be reduced because they
would be adsorbed onto the greater number of accumulation mode particles. At 7.0 bar
IMEP, it would be expected that the accumulation mode particles would be increased too,
resulting, again, in nucleation mode particle decreases, due to poorer mixture preparation
resulting from the greater KLMBT spark timing advances. At 8.5 bar IMEP, it is expected
that there would be no significant effect of the lower intake manifold temperature on the PM
117
Fig. 4-6a: PM (5.5 bar)
1.E+09
1.E+08
dN/dlogDp (#/cm3)
1.E+07
1.E+06
1.E+05
10 100 1000
Particle Diameter (nm)
0% EGR 4% EGR 8% EGR
12% EGR 13% EGR
1.E+08
dN/dlogDp (#/cm3)
1.E+07
1.E+06
1.E+05
10 100 1000
Particle Diameter (nm)
0% EGR 4% EGR 8% EGR 12% EGR
118
Fig. 4-6c: PM (8.5 bar)
1.E+09
1.E+08
dN/dlogDp (#/cm3)
1.E+07
1.E+06
1.E+05
10 100 1000
Particle Diameter (nm)
0% EGR 2% EGR 3% EGR
Figure 4-6 PN emissions versus particle diameter at KLMBT spark timings at a) 5.5 bar
IMEP, b) 7.0 bar IMEP and c) 8.5 bar IMEP
The NOx emission at the KLMBT spark timings for the three engine load conditions at the
tested EGR ratios is presented in Figure 4-7. The NOx emission increased significantly as the
engine load was increased due to the higher calculated average in-cylinder temperatures
which increased the formation of NO x in the hot flames. There is a clear trend of the NO x
decreasing as the EGR percentage was increased. EGR reduced the calculated average in-
cylinder temperatures (Figures 4-2a, 4-2b and 4-2c) for the three engine loads tested,
resulting in the NOx reductions observed. The overall reduction for the loads of 5.5, 7.0 and
8.5 bar IMEP are 64.3, 50.9 and 12.2%, respectively, showing the significant potential EGR
The rate of decrease of the NO x emission with EGR addition became greater as the engine
load was increased; however, the NOx decrease in terms of percent, over the achievable EGR
range for the respective loads, reduced as the engine load was increased. The rate of decrease
119
became greater as the engine load was increased because EGR addition was more effective at
reducing the average in-cylinder temperature as the engine load was increased. This can be
seen in Figures 4-2a, 4-2b and 4-2c, with similar calculated average in-cylinder temperature
reductions achieved for the three engine loads tested, despite the maximum EGR addition
reducing as the engine load was increased. It is thought that the greater mass of EGR gases as
the engine load was increased; at equivalent EGR additions in terms of percent, due to the
greater mass of intake air, caused the in-cylinder temperature to be suppressed more
significantly, resulting in the observed behaviour. It is thought that the NO x decrease in terms
of percent reduced as the engine load was increased, because of the reduced maximum EGR
At 5.5 bar IMEP, the NOx emission would have been reduced more significantly if the intake
manifold temperature increases could have been eliminated, due to the lower expected
combustion temperatures. At 7.0 and 8.5 bar IMEP, there would be no significant change
expected because, as discussed, the peak of the calculated average in-cylinder temperatures
would not have changed significantly. Again, it is thought that the improved KLMBT spark
timing would have quickly compensated for the decreased intake gas temperature by the time
temperatures reached the beginning of the range reported (500˚C). Thus, the effect on the
NOx emission which occurs at temperatures much higher than 500˚C would not have been
significant.
120
11
10
8
NOx (g/kWh)
2
0 2 4 6 8 10 12 14
EGR (%)
5.5 bar 7.0 bar 8.5 bar
Figure 4-7 NOx emissions versus EGR ratio at KLMBT spark timings
The HC emission data for the three tested engine loads across their respective EGR ranges is
shown in Figure 4-8. At 0% EGR, lower HCs were observed as the engine load was increased
because the calculated average in-cylinder temperatures were higher which increased the
oxidation rate of the HCs, reducing the numbers observed. The reduced calculated average
in-cylinder (Figures 4-2a-c) and exhaust (Figure 4-5) gas temperatures across the EGR range
for the three tested engine loads resulted in the HC emission increases observed across the
ranges for the respective loads. They increased by 48.3, 19.0 and 6.8% across the EGR range
for the engine loads of 5.5, 7.0 and 8.5 bar IMEP, respectively. It is proposed that the reduced
in-cylinder temperatures increased the thickness of the quench layers covering the
combustion chamber surface and slowed down and reduced the evaporation of impinged fuel,
resulting in the HC emission increase. The reduced exhaust temperatures reduced the
oxidation rate of HCs contributing to the increases observed. Some researchers have observed
similar overall HC increases with EGR addition (Alger et al., 2009a, Diana et al., 1996,
Takaki et al., 2014, Zhang et al., 2014b), but others have observed unchanged HC (Grandin et
121
The reduced time available for the fuel-air mixture preparation due to the KLMBT spark
advances across the EGR range will have resulted in poorer charge preparation, leading to
more pockets of a high local equivalence ratio, also contributing to the HC emission increases
observed. The same reasoning can also help explain why the HC emissions reduced as the
engine load was increased and the spark timing became more retarded.
There was a small but significant reduction in the rate of HC emission increase at 8% EGR
for 5.5 bar IMEP. This is mirrored by the slight decrease in the rate of reduction in NO x
emissions at the same condition (Figure 4-7). It is believed that the relatively smaller change
in the MFB profile (Figure 4-3a) between the EGR ratios of 4 and 8%, as compared to that
observed between the EGR ratios of 0 and 4%, led to a smaller reduction in the ROHR,
resulting in this behaviour. The HC emission decreased slightly for 5.5 bar IMEP between the
EGR additions of 12 and 13%. Although a decrease was observed it is not considered to be
significant, as can be seen from the confidence intervals on the figure. Therefore, it will not
be commented on further.
If the intake manifold temperature increases could have been eliminated, it is expected that
the HC emission reduction at 5.5 bar IMEP would have been increased, due to the higher
EGTs which could be expected in such a case. At 7.0 and 8.5 bar IMEP, the effect of the
lower EGTs would be expected to be cancalled out by the increased post-combustion time
provided by the greater KLMBT spark timing advance. Thus, it is not thought that the intake
manifold temperature increases with EGR additon had a significant effect on the HC
Despite the increases in HCs observed across the EGR range for the three engine loads tested,
it would be expected that a traditional TWC would be able to handle these increases to
maintain the vehicle-out HC emissions at non-EGR levels. This is based on the work of
122
(Hoffmeyer et al., 2009) who found that HCs are almost completely reduced by the catalyst,
even as the EGR rate is increased, due to an increased conversion efficiency of the catalyst.
However, more research would be required to confirm whether it can be achieved in this
engine.
11.5
10.5
9.5
HC (g/kWh)
8.5
7.5
6.5
5.5
4.5
0 2 4 6 8 10 12 14
EGR (%)
5.5 bar 7.0 bar 8.5 bar
123
4.3.2 Effect of Different EGR Types on Combustion and Gaseous Emissions
in a DISI Engine
The KLMBT spark timings, as shown in Table 4-4, could be advanced significantly with
EGR addition for all the EGR types tested. They were advanced by 10, 10, 12 and 14 CAD
for the hot EGR, cooled EGR, hot EGR after TWC and cooled EGR after TWC EGR types,
respectively. The EGR addition enabled the spark to be advanced significantly, through the
suppression of temperature rises in the end-zone. The KLMBT spark timing was most
advanced for the cooled EGR after TWC condition across the EGR range because the
removal of NO from the EGR gases reduced the knocking tendency of the engine (Kawabata
et al., 1999) and the cooler EGR gases reduced the temperature in the end-zone of the
combustion chamber. These reasons also explain the greater KLMBT spark timing advances
for hot EGR after TWC and cooled EGR, respectively, as compared to the hot EGR
condition.
Despite this, KLMBT spark timing advances would likely have been greater if intake
manifold temperature increases resulting from the EGR addition could have been eliminated,
due to the reduced end-zone temperatures. It must be noted that knocking could have been
avoided if the compression ratio was fully optimized for the fuel used, as mentioned
previously.
The EGR addition could only be increased to 10% for the cooled EGR condition because of
the pressure loss in the EGR heat exchanger. This means that a pump would be required to
achieve higher levels of EGR or a more efficient heat changer in terms of cooling power vs.
pressure loss would be required. A pump would reduce any fuel consumption benefits of the
cooled EGR system, therefore reducing any advantages it may have over the hot EGR
124
system. Despite the pressure loss in the heated line for the hot EGR after TWC condition, this
could be compensated for by increasing the gas flow rate of the compressed gas bottle,
Table 4-4 KLMBT spark timings (˚bTDC) (EGR addition & EGR type)
EGR Type
Hot EGR 14 18 24
Cooled EGR 20 24
HE after TWC 20 26
CE after TWC 20 28
The in-cylinder pressures for the four EGR types at the KLMBT spark timings, as shown in
Figures 4-9a, 4-9b, 4-9c and 4-9d, generally increased with EGR addition and the subsequent
spark advances which could be achieved, and the maximum in-cylinder pressure became
slightly earlier due to the more advanced spark timings. This was because the KLMBT spark
timing advances resulted in the heat being released from the fuel at an earlier CAD, enabling
higher in-cylinder pressures to be achieved before the downwards piston motion caused the
in-cylinder pressure to decrease again. However, the maximum in-cylinder pressure remained
approximately the same across the EGR range for hot EGR (Figure 4-9a), with the peak
pressure becoming less advanced than with the other EGR types. This was due to the lower
spark advance achieved at equivalent EGR ratios (Table 4-4) and the resulting retardation of
Greater in-cylinder pressure increases would have been likely if the intake manifold
temperature increases, resulting from the EGR addition, could have been eliminated. This is
because it would likely have led to greater KLMBT spark timing advances, as discussed.
125
Fig. 4-9a: In-cylinder pressure (hot EGR)
40
30
25
20
15
10
0
-50 0 CAD 50 100
0% EGR 7% EGR 14% EGR
40
35
30
25
20
15
10
0
-50 0 CAD 50 100
0% EGR 7% EGR 10% EGR
126
Fig. 4-9c: In-cylinder Pressure (Hot EGR after TWC)
35
30
25
20
15
10
0
-50 0 CAD 50 100
0% EGR 7% EGR 14% EGR
35
30
25
20
15
10
0
-50 0 CAD 50 100
Figure 4-9 In-cylinder pressures versus CAD at KLMBT spark timings for a) hot EGR,
b) cooled EGR, c) hot EGR after TWC and d) cooled EGR after TWC
The calculated average in-cylinder temperatures, as shown in Figures 4-10a, 4-10b, 4-10c and
4-10d at the KLMBT spark timings, decreased across the EGR range for the different EGR
types due to the temperature suppression effect of the EGR gases. The temperatures decreases
across the EGR range were greater for the cooled EGR and cooled EGR after TWC EGR
127
types as compared to the hot EGR and hot EGR after TWC EGR types due to their lower
It is believed that the increased intake manifold temperatures resulting from the EGR addition
were compensated for by the less optimized combustion phasing resulting from the
consequently reduced KLMBT spark timing advance (Table 4-4), compared to the maximum
spark advance possible, as explained previously. It is thought that the improved KLMBT
spark timing would have quickly compensated for the intake gas temperature decrease by the
time temperatures reached the beginning of the range reported (500˚C); thus, the effect would
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
0% EGR 7% EGR 14% EGR
128
Fig. 4-10b: In-cylinder Temperature (cooled EGR)
1900
1500
1300
1100
900
700
500
-50 0 CAD 50 100
0% EGR 7% EGR 10% EGR
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
0% EGR 7% EGR 14% EGR
129
Fig. 4-10d: In-cylinder Temperature (cooled EGR after TWC)
1900
1500
1300
1100
900
700
500
-50 0 CAD 50 100
0% EGR 7% EGR 14% EGR
Figure 4-10 Calculated average in-cylinder temperatures versus CAD at KLMBT spark
timings for a) hot EGR, b) cooled EGR, c) hot EGR after TWC and d) cooled
EGR after TWC
4.3.2.3 MFB
The MFB profiles at the KLMBT spark timings, as shown in Figures 4-11a, 4-11b, 4-11c and
4-11d for the four EGR types, started earlier and became elongated as the EGR ratio was
increased, due to the spark timing becoming more advanced and due to the laminar flame
speed reducing effects of the EGR gases, respectively (Rhodes and Keck, 1985). The MFB
profile for the cooled EGR condition (Figure 4-11b) was approximately the same at the EGR
ratios of 7 and 10%. This is because the rate of spark advance between 7 and 10% EGR was
1.33 CAD/EGR (%), which was higher than the 0.86 CAD/EGR (%) rate of spark advance
achieved between the EGR ratios of 0 and 7%. This allowed the MFB profile to be
maintained with little change between 7 and 10% EGR addition whereas between 0 and 7%
EGR addition, the less significant spark advance caused the MFB profile to become more
elongated, producing the observed behaviour. For a quantitative analysis of the combustion
130
speed, the MFB50 and MFB10-90 have been calculated from the MFB profiles (please see
next section).
Again, KLMBT spark timing advances would likely have been greater if intake manifold
temperature increases, resulting from the EGR addition, could have been eliminated; thus, the
MFB profiles would likely have been advanced slightly in comparison to those in Figures 4-
11a-d. The profile would also become slightly more elongated because the laminar flame
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
0% EGR 7% EGR 14% EGR
131
Fig. 4-11b: MFB (cooled EGR)
1.0
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
0% EGR 7% EGR 10% EGR
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
0% EGR 7% EGR 14% EGR
132
Fig. 4-11d: MFB (cooled EGR after TWC)
1.0
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
0% EGR 7% EGR 14% EGR
Figure 4-11 MFB versus CAD at KLMBT spark timings for a) hot EGR, b) cooled EGR,
c) hot EGR after TWC and d) cooled EGR after TWC
The MFB50 at the KLMBT spark timings became earlier across the EGR range for the three
conditions of cooled EGR, hot EGR after TWC and cooled EGR after TWC as shown in
Figure 4-12. This is because, as discussed, the KLMBT spark timings (Table 4-4) could be
advanced significantly with the EGR addition. However, for the hot EGR condition, the
MFB50 became later. It seems therefore that the KLMBT spark timing advance achieved was
not sufficient to overcome the laminar flame speed slowing effects of the EGR addition,
whereas for the other EGR types, it was. This is because the spark timing advance was less
significant for the hot EGR condition as compared to the other EGR conditions, because it
did not benefit from the cooled EGR gas or from the removed NO; the other three EGR types
benefited from at least one of these things. Therefore, the lesser spark advance was not able
to overcome the combustion slowing effects of the EGR gases, but the greater spark advances
achieved with the other EGR types were sufficient to overcome them.
133
Despite the same KLMBT spark timings being achieved for the cooled EGR, hot EGR after
TWC and cooled EGR after TWC at 7% EGR addition, the MFB50 data is not the same. It is
proposed that this is due to the different intake plenum temperatures (Figure 4-18) which
meant that the fuel-air charges were of different temperatures before combustion, resulting in
a variation in flame speeds and thus slightly different MFB50 data, despite the same spark
timings.
Again, KLMBT spark timing advances would likely have been greater if intake manifold
temperature increases, resulting from the EGR addition, could have been eliminated. Thus,
18.5
18.0
17.5
MFB50 (˚aTDC)
17.0
16.5
16.0
15.5
15.0
14.5
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
The combustion duration (MFB10-90) at the KLMBT spark timings, as shown in Figure 4-
13, increased significantly for all of the EGR types across the EGR range because of the
aforementioned laminar flame speed reducing effects of the EGR gases (Rhodes and Keck,
1985). The cooled EGR and cooled EGR after TWC EGR types were more significantly
affected than the hot EGR and hot EGR after TWC EGR types. It is proposed that this is
134
because the flame travelled at a slower speed in the cooler fuel-air charge. The cases with the
simulated TWC were more greatly affected then their corresponding hot and cooled
conditions respectively. It is proposed that this was because of the slightly reduced intake
plenum temperatures (Figure 4-18) for the TWC conditions as compared to their
corresponding hot and cooled conditions. This is thought to have reduced the laminar flame
speed which increased the combustion duration for these in respect to their corresponding
non-TWC conditions.
As explained, if the intake manifold temperature increases could have been eliminated, it is
believed that the flame would travel more slowly in the resulting cooler fuel-air charge,
increasing the combustion duration slightly. If the compression ratio was optimised for the
fuel used, the combustion phasing would have been optimised for each test point, therefore, it
would be expected that the combustion at all of the conditions tested would have been
31
30
29
28
MFB10-90 (CAD)
27
26
25
24
23
22
21
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
135
4.3.2.5 Fuel Consumption
Gravimetric ISFCnet at the KLMBT spark timings decreased for all EGR types as shown in
Figure 4-14. The fuel consumption reductions were due to MFB50 advances for the cooled
EGR, hot EGR after TWC and cooled EGR after TWC EGR types, along with pumping loss
explained in the first part of Chapter 4. It must be noted that if the hot and cooled EGR after
TWC EGR types had used a real TWC rather than a simulated TWC then their fuel
consumption would have been higher across the EGR range due to the increased back-
Overall, the fuel consumption decreased by 1.8, 3.8, 1.6 and 2.9%, respectively, across the
EGR range, for the hot EGR, cooled EGR, hot EGR after TWC and cooled EGR after TWC
EGR types. Despite these fuel consumption reductions observed, it must be noted that the
VAF had an average error of 1.6%, as discussed, therefore, the reduction of 1.6% observed
across the EGR range for the hot EGR after TWC condition is limited, because it could have
resulted from this error. The reductions for the other EGR types are however not considered
to be limited. Again, KLMBT spark timing advances would likely have been greater if intake
manifold temperature increases, resulting from the EGR addition, could have been
eliminated. Thus, the fuel consumption improvements would likely have been greater.
The fuel consumption reduction was greatest for the cooled EGR condition despite the
limited EGR that could be achieved, because its MFB50 (Figure 4-12) was further advanced
by the EGR addition and consequent spark advances than that of the other EGR types. In
addition, its combustion duration (Figure 4-13) did not increase as significantly as that of the
cooled EGR after TWC EGR type, improving the conversion efficiency of heat and pressure
into piston work through improved combustion phasing. The next greatest fuel consumption
136
reduction achieved was for cooled EGR after TWC. This is because the KLMBT spark
timing advance (Table 4-4) was the greatest, therefore allowing the MFB50 to be advanced
closer to its optimum phase (Figure 4-12). Fuel consumption reductions were achieved with
hot EGR despite its MFB50 becoming retarded with EGR addition (Figure 4-12). This is
because its PMEP was significantly reduced across the EGR range (Figure 4-15) and it
efficiency.
The hot EGR after TWC EGR type achieved fuel consumption improvements of lesser
magnitude than the cooled EGR and cooled EGR after TWC conditions because its KLMBT
spark timing could not be advanced as much, meaning its MFB50 (Figure 4-12) was more
retarded than those of the cooled EGR and cooled EGR after TWC EGR types. At 7% EGR,
its MFB50 was more advanced than that of the cooled EGR after TWC EGR type (Figure 4-
12) and its PMEP was lower (Figure 4-15), but its fuel consumption was still higher. It is
proposed that this was due to the higher peak of the calculated average in-cylinder
temperature (1844˚C) compared to that of the cooled EGR after TWC EGR type (1823˚C), as
shown in Figures 4-10c and 4-10d, respectively. It is thought that this increased heat transfer
to the combustion chamber walls, reducing the thermal efficiency, thus causing the fuel
consumption to be higher.
Hot EGR had lower fuel consumption at the maximum EGR addition of 14% than the hot
EGR after TWC EGR type. Despite this observation, it is not considered to be significant
considering it is within the margin of error, as indicated by the relevant confidence intervals
(Figure 4-14). For the cooled EGR and cooled EGR after TWC EGR types, the fuel
consumption reduction became less significant as the EGR addition was increased from 7%
to its maximum, as compared to that achieved as the EGR addition was increased from 0 to
137
7%. This observation is also not considered to be significant because it is within the margins
If the intake plenum temperature increases (Figure 4-18) could have been eliminated, it is
believed that the fuel consumption improvements would have been greater, especially for hot
EGR, due to the greater KLMBT spark timing advances and thus more optimized MFB50
260
258
256
ISFCnet (g/kWh)
254
252
250
248
246
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
Figure 4-14 Gravimetric ISFCnet versus EGR ratio at KLMBT spark timings
4.3.2.6 PMEP
The PMEP at the KLMBT spark timings, as shown in Figure 4-15, decreased for all of the
EGR types across the EGR range, because as mentioned, the throttle valve had to be opened
more to provide the same amount of fresh air as in the non-EGR case, or in other words, the
It is proposed that the greatest reduction across the EGR range was achieved with the hot
EGR condition because the intake plenum temperature increased the most (Figure 4-18). This
138
caused the intake air to become less dense meaning that the throttle had to be more opened in
order to maintain a stoichiometric mixture than it did with the other EGR types, thus causing
the pumping losses to reduce more significantly. For hot and cooled EGR, the reduction was
greatest as the EGR was increased from 7% to its maximum, whereas the greatest reduction
for hot and cooled EGR after TWC occurred as the EGR was increased from 0 to 7%. It is
thought that the more significant increase in intake plenum temperature (Figure 4-18) for hot
EGR, as the EGR addition was increased from 7 to 14%, meant that the throttle had to be
opened more than when the EGR addition was increased from 0 to 7%, resulting in the
observed behaviour. The observed behaviour for the cooled EGR condition is within the
significant.
Pumping work was greater for the cooled EGR after TWC EGR type than the hot EGR after
TWC EGR type. It is proposed that this is due to its lower intake plenum temperature (Figure
4-18). This meant that the intake air was denser so the throttle had to be opened less to allow
the same amount of fresh air into the combustion chamber as with the hot EGR after TWC
EGR type, thus increasing the pumping work in comparison. It must be noted that if the hot
and cooled EGR after TWC EGR types had used a real TWC rather than a simulated TWC,
their PMEP would have been greater across the EGR range due to the increased back-
The PMEP reduction would likely have been less significant if intake manifold temperature
increases, resulting from the EGR addition, could have been eliminated. This is because the
intake air would have become denser, necessitating the throttle to be more closed to allow the
same amount of fresh air into the combustion chamber, thus increasing the pumping work.
139
0.35
0.33
0.31
PMEP (bar)
0.29
0.27
0.25
0.23
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
4.3.2.7 EGT
The EGT at the KLMBT spark timings decreased significantly for all of the EGR types
across the EGR range, as shown in Figure 4-16. Overall, it decreased by 40.4, 41.6, 46.9 and
52.1˚C for the EGR types of hot EGR, cooled EGR, hot EGR after TWC and cooled EGR
after TWC, respectively. This is because the KLMBT spark timing advances (Table 4-4)
resulted in the pressure and heat from combustion being more efficiently converted into
The temperatures reduced for hot EGR despite its MFB50 becoming later. It is proposed that
the earlier KLMBT spark timing (Table 4-4) resulted in the energy from the fuel being
released earlier. This resulted in energy being more efficiently converted from the
combustion in piston work, despite the slightly later MFB50 (Figure 4-12) and the longer
combustion duration (Figure 4-13) across the EGR range, resulting in reduced EGTs across
140
At 10% EGR addition, the greatest reductions were observed for cooled EGR. It is thought
that this is because its MFB50 (Figure 4-12) was the most advanced at 10% EGR addition,
meaning that the pressure and heat from the combustion was more efficiently converted into
piston work. This caused the EGT to decrease more than with the other EGR types which had
a later MFB50 at 10% EGR addition. Cooled EGR after TWC then hot EGR after TWC had
the next greatest rate of reduction across their tested EGR ranges. Again, this can be
explained by the more advanced MFB50 for the cooled EGR after TWC EGR type as
compared to the hot EGR after TWC EGR type (Figure 4-12).
Again, KLMBT spark timing advances would likely have been greater if intake manifold
temperature increases, resulting from the EGR addition, could have been eliminated. This
would have likely resulted in greater EGTs decreases across the EGR range, as discussed
previously.
660
650
640
EGT (˚C)
630
620
610
600
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
141
4.3.2.8 COVIMEP
The COVIMEP at the KLMBT spark timings increased significantly across the EGR range for
all of the EGR types, as shown in Figure 4-17. It is proposed that this is because the
combustion speed was reduced across the EGR range (Figure 4-13) which meant that the
flame was more affected by the airflow in the combustion chamber, thus increasing the
COVIMEP observed. As can be seen from the confidence intervals indicated on the figures, the
differences between the different EGR types could have resulted from experimental error,
As discussed, it is believed that the combustion duration would have been slightly increased
if the intake manifold temperature increases could have been eliminated. This would be
expected to increase the COVIMEP slightly because the flame would be more affected by the
4.4
3.9
COVIMEP (%)
3.4
2.9
2.4
1.9
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
142
4.3.2.9 Intake Plenum Temperature
The intake plenum temperature at the KLMBT spark timings, as shown in Figure 4-18,
increased significantly across the EGR range for all of the EGR types. This was expected for
the hot EGR and hot EGR after TWC EGR types but it was less expected for the cooled EGR
and cooled EGR after TWC EGR types. It is proposed that the cooled EGR gases were not as
cold as the engine fresh intake air, therefore causing the intake plenum temperatures to
increase. The intake temperatures are still lower for the cooled EGR types as compared to the
hot EGR types, as is to be expected. Any further relative differences between the different
EGR types are not considered to be significant as can be seen from the confidence intervals
38
36
Intake Plenum Temperature (˚C)
34
32
30
28
26
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
Figure 4-18 Intake plenum temperature versus EGR ratio at KLMBT spark timings
143
4.3.2.10 Gaseous Emissions (NOx and HC)
The NOx emission at the KLMBT spark timings, as shown in Figure 4-19, decreased
significantly by 50.5, 31.3, 43.7 and 45.3% across the EGR range for the hot EGR, cooled
EGR, hot EGR after TWC and cooled EGR after TWC EGR types, respectively. It is
proposed that the reduced calculated average in-cylinder temperatures across the EGR range
(Figures 4-10a-d) resulted in a reduced oxidation rate of in-cylinder nitrogen gas in the hot
The reduction was greatest for hot EGR, similar for hot and cooled EGR after TWC, and it
was the least for cooled EGR. It is thought that this is because the MFB50 (Figure 4-12)
became later for hot EGR across the EGR range. The MFB50 for cooled and hot EGR after
TWC were similar across the EGR range therefore explaining why their NOx emission was
similar, while the MFB50 was the most advanced for the cooled EGR condition, resulting in
the highest NOx emission across the EGR range. With an advanced MFB50, the fuel-air
mixture burns at an earlier point in the engine cycle where the in-cylinder turbulence is
stronger, causing the combustion to proceed more quickly. It is thought that this resulted in
higher local temperatures which resulted in the differences in NOx emissions between the
different EGR types across the EGR range. While there were differences in the peaks of the
calculated average in-cylinder temperatures for the different EGR types (Figure 4-10a-d), it is
thought that the differences in MFB50 and their subsequent effects on local in-cylinder
temperatures were more significant regarding the NOx behaviour. This is because the average
As discussed, it is not believed that the intake manifold temperature increases with EGR
addition had a significant effect on the peak calculated average in-cylinder temperatures.
144
Therefore, it is not thought that the NOx emission was significantly affected by the intake
8.5
8.0
7.5
7.0
NOx (g/kWh)
6.5
6.0
5.5
5.0
4.5
4.0
3.5
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
Figure 4-19 NOx emissions versus EGR ratio at KLMBT spark timings
The HC emission at the KLMBT spark timings significantly increased across the EGR range
for all of the EGR types tested, as shown in Figure 4-20. It was increased by 14.7, 11.0, 12.9
and 10.1% for the four respected EGR types of hot EGR, cooled EGR, hot EGR after TWC
and cooled EGR after TWC. It is thought that the in-cylinder temperature (Figures 4-10a-d)
and EGT decreases across the EGR range (Figure 4-16) reduced the oxidation of HCs for all
of the EGR types in the engine cylinder and exhaust, respectively, leading to the HC
increases across the EGR range. Also, the increased combustion duration (Figure 4-13)
reduced the time available for post-combustion oxidization of HCs for all of the EGR types,
The increase was greatest for hot EGR, followed by hot EGR after TWC, then cooled EGR,
then cooled EGR after TWC. It is proposed that the later MFB50 for hot EGR and hot EGR
after TWC across the EGR range (Figure 4-12) reduced the time for post-combustion
145
oxidation, causing their HC emission to increase the most. Although the MFB50 for the
cooled EGR condition was more advanced than that of the hot and cooled EGR after TWC
conditions, its HC emission across its EGR range of 0 to 10% was similar to that of the hot
EGR after TWC condition and higher than that of the cooled EGR after TWC condition
within the same 0 to 10% EGR range. This is because it introduced additional HCs into the
combustion chamber while the hot and cooled EGR after TWC conditions did not, because
the simulated EGR gas was taken from the gas bottle. Therefore, the additional HCs reduced
the oxidation of the HCs in the combustion chamber producing the higher than expected HC
emission. With a real TWC, most but not all of the HCs would have been removed from the
EGR gas, resulting in a similar effect, although its significance would be reduced. It is
proposed that the longer combustion duration for the cooled EGR after TWC condition as
compared to the hot EGR after TWC condition (Figure 4-13) resulted in an improvement in
As discussed in the first part of this chapter, it is not believed that the intake manifold
temperature increases had a significant effect on the HC emissions observed. Again, despite
the increases in HCs observed across the EGR range for the four EGR types tested, it would
be expected that a traditional TWC would be able to handle this increase to maintain the
146
7.4
7.2
7.0
HC (g/kWh)
6.8
6.6
6.4
6.2
0 2 4 6 8 10 12 14 16
EGR (%)
Hot EGR Cooled EGR Hot EGR after TWC Cooled EGR after TWC
147
4.5 Conclusions
The effect of hot EGR on the combustion and emissions of a single cylinder DISI research
engine was investigated in this chapter and the following conclusions have been made.
1. KLMBT spark timing was advanced significantly by 18, 8, and 4 CAD for the three
respective engine loads of 5.5, 7.0 and 8.5 bar IMEP, due to EGR suppressing the
2. Fuel consumption was reduced by 2.2, 4.1, and 1.0% for the three respective engine loads
with EGR addition, because the pumping work was reduced for all engine loads and the
KLMBT could be advanced towards its optimum phasing for the engine loads of 7.0 and
3. EGR addition increased the accumulation mode particles and reduced the nucleation
mode particles for the loads of 5.5 bar IMEP (from 0 to 12% EGR addition) and 7.0 bar
IMEP, because of the combustion temperature suppression effect of the EGR, and
However, the nucleation mode particle reduction for 7.0 bar IMEP was not significant
and the accumulation mode particles reduced as the EGR was increased from 12 to 13%
for the load of 5.5 bar IMEP. In addition, there was not a significant change in nucleation
and accumulation mode particles for the load of 8.5 bar IMEP because of the low 3%
These main findings from this section are also summarized in Table 4-5.
148
Table 4-5 Results summary (EGR addition & engine load) (highlighted=improvement,
underlined=worsening)
8.5 4 CAD 1.0% 7.6˚C No sig. change No sig. change 12.2% 6.8%
The effect of different EGR types of hot EGR, cooled EGR, HE after TWC and CE after
TWC on the combustion and gaseous emissions of a single cylinder DISI research engine was
also investigated in this chapter and the following conclusions have been made.
4. KLMBT spark timing was advanced significantly by 10, 10, 12 and 14 CAD,
respectively, for the four EGR types, due to EGR suppressing the temperature rises in the
combustion chamber end-zone; cooler EGR gases and the removal the NO from the EGR
gases accounted for the differences between the different EGR types.
5. Fuel consumption was reduced by 1.8, 3.8, 1.6 and 2.9% for the four respective EGR
types, because pumping losses were reduced for all EGR types and the MFB50 could be
6. The NOx emission was reduced significantly by 50.5, 31.3, 43.7 and 45.3% across the
EGR range for the four respective EGR types, due to the temperature suppression effect
of the EGR gases. MFB50 combustion phasing differences resulting from the different
KLMBT spark timing advances achieved accounted for the differences between the
149
7. Overall, if fuel consumption is the priority than the cooled EGR type is the best EGR
type, while if emission reduction (NOx and HC) is the priority than cooled EGR after
The main findings from this section are also summarized in Table 4-6.
Table 4-6 Results summary (EGR addition & EGR type) (highlighted=improvement,
underlined=worsening)
EGR Type
Hot EGR 10 CAD 1.8% 40.4˚C 50.5% 14.7%
150
Chapter 5
The Effect of Intake Airflow and Hot EGR on Engine Combustion and PM
Emissions
The aim of this chapter is to provide details of the combustion and PM emissions
characteristics of a DISI engine operated with swirl and tumble intake airflows, and with
5.1 Introduction
The main combustion parameters investigated in this chapter are KLMBT spark timing, in-
cylinder pressure, calculated average in-cylinder temperature, MFB, fuel consumption and
There is an ever growing demand for reduction of PM emissions from engines, as emission
regulations become stricter, particularly with the Euro VI regulation which has limited both
particulate mass and size for the first time in DISI engines. It is well known that swirl and
tumble have a significant effect on air-fuel mixture formation as well as on the combustion
and emissions of IC engines. However, what is not well known is the effect they have on the
formation of particulates in DISI engines, and what effect combining swirl and tumble with
EGR has on the combustion and PM emissions. EGR suppresses the temperature rise in the
combustion chamber, reducing NO x emissions and allowing the KLMBT spark timing to be
advanced, as well as enabling the throttle to be more opened, providing significant fuel
economy benefits. However, swirl and tumble affect the speed of the combustion, which
affects the laminar speed slowing effect of the EGR gases. Therefore, it is important to
understand the influence of swirl and tumble, combined with EGR, on combustion and PM
emissions. The Euro VI emissions regulation has increased interest in the role of EGR
151
addition on engine particulates, with the potential of EGR coming under question because of
it has been observed to have a negative effect on particulates (Gill et al., 2011, Ma et al.,
Overall, the author recognizes an opportunity to extend the research of swirl and tumble in
DISI engines further by investigating the effect of swirl and tumble, and EGR addition on PM
emissions in this chapter. By doing this, it is hoped that the effect can be quantified in order
to inform the engine design process in terms of the selection of swirl and tumble ratios, and
the possible incorporation of a swirl/tumble valve into DISI engines to help achieve these at
specific engine speeds and loads. The author also recognizes that the Euro VI emissions
regulation has led many to question the potential of EGR, considering it is generally
considered to increase engine particulates. Thus, this is another area researched further in this
chapter.
The experimental test procedure outlined in Section 3.9 of this thesis was followed in order to
conduct the engine tests in the single-cylinder engine to collect the data presented and
discussed in this chapter. Relative air-fuel ratio λ was maintained at 1 during the experiments
and a COVIMEP of 5% was not exceeded. Valve timings were set at IVO=16˚bTDC and
EVC=36˚aTDC. KLMBT spark timings and a fixed geometric compression ratio of 11.5
were used. The baffle plates as shown and discussed in Section 3.2.7 were used to create the
swirl and high tumble intake airflows in the engine, in addition to what is already produced
by the relevant engine geometry, such as the design of the intake runner. As well as these two
plates, another baffle plate was made with the same profile as the intake runner, so that a
reference low tumble condition could be tested with the same intake geometry as the swirl
152
and high tumble conditions. The estimated swirl/tumble ratios for the three baffle plates are
provided in Table 3-2. In order to achieve hot EGR in the engine, and in order to achieve the
desired engine load and EGR addition, the equipment setup and test procedure outlined in
The load of 7.0 bar IMEP was chosen to study because it represents a medium-high load in
this engine, and while it produces knock, the knock produced is of a lower magnitude than
that at higher engine loads, allowing swirl and tumble to be studied without causing any
significant engine damage. It also allowed a reasonable, while not a high EGR addition of
14% to be achieved. A higher engine load would have meant that the throttle would have
needed to be more opened, thus reducing the maximum level of EGR that could have been
achieved. One load was chosen to study so that an in-depth understanding could be formed,
which would not be possible if multiple loads were studied with the different intake airflows.
The spark was swept for each intake airflow condition and EGR ratio in order to find the
KLMBT spark timing using the MatLab script pressure trace analysis technique outlined in
Section 3.6.9. A research grade gasoline was used to obtain the results in this chapter with the
properties outlined in Section 3.8. The SMPS 3936 was used to obtain the PM emission
results in this chapter. Gaseous emissions were not measured in this chapter because it is
The test matrix provided in Table 5-1 was carried out in order to investigate the intake
airflow and EGR effect on engine combustion and PM emissions at the single engine load of
153
Table 5-1 Experiment test matrix (intake airflow & EGR addition)
Intake
Air-Flow
Low Tumble 1 2 3 4
Swirl 5 6 7 8
High Tumble 9 10 11
From the KLMBT spark timings, as shown in Table 5-2 for the different intake airflow
conditions, it can be seen that the spark was retarded for the swirl and high tumble conditions,
as compared to the low tumble condition; the high tumble condition being the most retarded.
Overall, the increased turbulence produced in the swirl and high tumble conditions (CID was
reduced by ~1 CAD with greater reductions observed with EGR addition) as compared to the
baseline low tumble condition meant that the spark timing had to be retarded in order for the
turbulence and thus combustion speed to decay in order to avoid knock. Although the
additional turbulence was allowed to dissipate somewhat by retarding the spark timing, it is
still believed to have improved mixture formation more significantly than spark retardation
alone would have done, which will be explored further in the emissions section.
From the MFB data (Figure 5-3a) it can be seen that the combustion duration was longer for
the swirl condition than it was for the low tumble condition (MFB was ~24.0 CAD for the
swirl condition and it was ~22.5 CAD for the low tumble condition, at 0% EGR addition).
Therefore, it is thought that this provided more time for the temperature to rise in the end-
zone of the combustion chamber, producing auto-ignition, resulting in the retarded KLMBT
154
spark timing for the swirl condition. It must be noted that knocking could have been avoided
if the compression ratio was fully optimized for the fuel used, as discussed previously.
For the high tumble condition, it is proposed that the velocity of the intake air was increased
significantly, causing the fuel-air mixture to burn much more quickly (MFB10-90 was ~21.0
CAD, compared to ~22.5 and ~24.0 CAD for the low tumble and swirl conditions,
respectively, at 0% EGR addition). It is believed that this increased the amount of heat
radiated to the fuel-air mixture in the end-zone of the combustion chamber which caused
knock to be initiated at an earlier spark timing than for the other two intake airflow
proposed that this effect was more significant than the faster flame speed, which does not
appear to have been fast enough to consume the auto-ignition site before it auto-ignited. This
evident relationship between flame speed and knock tendency should be investigated further
The end-zone temperatures were reduced (allowing the calculated average in-cylinder
temperatures to be maintained or their reductions minimized (Figures 5-2b, 5-2c and 5-2d))
by the combustion slowing effects of the EGR addition (Rhodes and Keck, 1985). This
allowed the spark timing to be advanced to an earlier point in the engine cycle where the in-
cylinder turbulence was stronger, cancelling out the effect EGR had on the laminar flame
At the high tumble condition with 0% EGR addition, the combustion was faster than with the
other two intake airflow conditions, as mentioned. This meant that good KLMBT spark
timing improvements could be made with EGR addition. Although EGR addition will have
allowed the KLMBT spark timing to be advanced at the low tumble and swirl conditions too,
155
it is proposed that the effect was greater with the high tumble condition. This is because its
KLMBT spark timings started further away from the optimal condition which meant that
greater (or equal) improvements in the KLMBT spark timing could be made at equivalent
EGR additions.
KLMBT spark timing advances would have been greater if the intake manifold temperature
increases resulting from the EGR addition could have been eliminated, as discussed
previously.
Table 5-2 KLMBT spark timings (˚bTDC) (intake airflow & EGR addition)
Intake
Air-Flow
Low Tumble 16 20 24 26
Swirl 8 10 12 14
High Tumble -2 2 6
The in-cylinder pressure data at 0% EGR addition presented in Figure 5-1a shows that the in-
cylinder pressure significantly decreased with the swirl and high tumble conditions, as
compared to the low tumble condition. This was due to the significant spark retard required
(Table 5-2), causing more energy to be released from the fuel during the expansion stroke,
once the in-cylinder pressure was already reducing. Therefore, this energy did not contribute
to the in-cylinder pressure increases, resulting in the in-cylinder pressure decreases observed.
The in-cylinder pressures across the EGR ranges, as presented in Figures 5-1b, 5-1c and 5-1d,
for the low tumble, swirl and high tumble conditions, respectively, generally increased across
the EGR range. The reason for the increases observed are the advances in the KLMBT spark
156
timing achieved with the EGR addition, releasing more of the energy from the fuel at an
earlier point in the engine cycle where it could contribute to the in-cylinder pressure
increases. However, for the maximum EGR addition of 11% at the low tumble and swirl
conditions, the in-cylinder pressure reduced. This is because the KLMBT spark timing
advance achieved with EGR addition was not sufficient at these conditions to overcome the
combusiton slowing effects of the additional EGR gases (Rhodes and Keck, 1985).
In-cylinder pressure increases with EGR addition will likely have been greater if the intake
manifold temperature increases could have been eliminated, due to the greater KLMBT spark
40
35
30
25
20
15
10
0
-50 0 CAD 50 100
Low Tumble Swirl High Tumble
157
Fig. 5-1b: In-cylinder Pressure (low tumble)
45
35
30
25
20
15
10
0
-50 0 CAD 50 100
40
35
30
25
20
15
10
0
-50 0 CAD 50 100
0% EGR 4% EGR 8% EGR 11% EGR
158
Fig. 5-1d: In-cylinder Pressure (high tumble)
45
35
30
25
20
15
10
0
-50 0 CAD 50 100
Figure 5-1 In-cylinder pressure verses CAD at KLMBT spark timings for a) the three
intake airflow conditions at 0% EGR, b) low tumble, c) swirl, d) high tumble
2a, was approximately the same for the swirl condition as it was for the low tumble condition,
although it was slightly retarded in comparison. The KLMBT spark timing had to be retarded
for the swirl condition as compared to the low tumble condition (Table 5-2), resulting in the
observed behaviour. For the high tumble condition, the calculated average in-cylinder
temperature significantly increased compared to the other two intake airflow conditions, due
to the faster combustion, as discussed previously. This significantly increased the rate that
heat was released from the fuel, resulting in the increased calculated average in-cylinder
temperature. The calculated average in-cylinder temperature profile for the high tumble
condition was also retarded in comparison, because of its KLMBT spark timing retard (Table
5-2).
There was no significant change in the magnitude of the calculated average in-cylinder
temperatures across the EGR range for the low and high tumble conditions as shown in
159
Figures 5-2b and 5-2d. As discussed, the EGR addition allowed the KLMBT spark timing to
the combustion temperature reducing effects of the EGR gases. They have a higher heat
capacity than the fresh air they replaced in the engine cylinder thus without advancing the
spark timing, the calculated average in-cylinder temperatures would have decreased.
However, the spark advance allowed them to be maintained across the EGR range by
releasing the heat earlier from the fuel. At 4 and 8% EGR addition for the low tumble intake
airflow it appears that the KLMBT spark timing advance was sufficient to increase the
conditions. For the swirl condition, as shown in Figure 5-2c, the KLMBT spark timing
advances were not sufficient to maintain the calculated average in-cylinder temperatures, thus
It is believed that the increased intake manifold temperatures resulting from the EGR addition
were compensated for by the less optimized combustion phasing resulting from the
consequently reduced KLMBT spark timing advance (Table 5-2), compared to the maximum
spark advance possible, as discussed in the first part of Chapter 4. Thus, it is thought the two
effects of increased intake manifold temperature and reduced subsequent KLMBT spark
timing advances cancelled one another out, meaning there was no overall effect on the peak
160
Fig. 5-2a: In-cylinder Temperature (three airflow conditions)
2100
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
1900
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
161
Fig. 5-2c: In-cylinder Temperature (swirl)
2100
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
1900
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
Figure 5-2 Calculated average in-cylinder temperature verses CAD at KLMBT spark
timings for a) the three intake airflow conditions at 0% EGR, b) low tumble,
c) swirl, d) high tumble
162
5.3.3 MFB
The low tumble and swirl intake airflow conditions had similar MFB profiles, as shown in
Figure 5-3a at 0% EGR addition, despite the retarded spark timing for the swirl condition as
compared to the low tumble condition. The swirl condition increased the speed of the initial
flame propagation compared to the low tumble condition by approximately 1 CAD, causing
the MFB50s at 0% EGR to be very similar. (Nagayama et al., 1977) also observed a similar
effect. However, the later stage of combustion appears to have been longer for the swirl
condition. This is most likely due to the breakdown of the swirl flow into weak general in-
cylinder turbulence. The high tumble condition in comparison had a significantly retarded
MFB profile compared to the low tumble and swirl conditions, despite its initial flame
propagation also being quicker. This was due to its retarded KLMBT spark timing (Table 5-
2). Unlike the swirl condition, it does not appear to suffer from the longer late stage of
combustion.
The MFB data, presented in Figures 5-3b, 5-3c and 5-3d for the three respected intake airflow
conditions of low tumble, swirl and high tumble, shows that the MFB profile generally
became advanced with EGR addition. This is because the KLMBT spark timing could be
advanced (Table 5-2), as explained in the first part of Chapter 4. It can be seen in Figure 5-3b
for the low tumble condition that 11% EGR addition resulted in a significantly increased
combustion duration. Therefore, it appears that the KLMBT spark advance achieved was not
sufficient to overcome the combustion slowing effects of the additional EGR, at this
condition. A similar effect occurred with the swirl intake airflow condition, but rather than
increasing the overall combustion duration, it produced a significantly slower earlier stage of
combustion than observed at lower EGR additions and a similar overall combustion duration.
163
Therefore, the additional EGR appears to have had a significant effect on this early stage of
Again, KLMBT spark timing advances would likely have been greater if intake manifold
temperature increases, resulting from the EGR addition, could have been eliminated. Thus,
the MFB profiles would likely have been advanced slightly in comparison to those shown in
Fig. 5-3b-d. In addition, it is expected that the flame would spread slightly more slowly in the
cooler charge, thus, the MFB profile would become slightly more elongated. If the
compression ratio was optimised for the fuel used, the combustion phasing would have been
optimised for each test point, therefore, it would be expected that the combustion of gasoline
would have been quicker due to higher in-cylinder turbulence at the time of ignition.
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0
-50 0 CAD 50 100
Low Tumble Swirl High Tumble
164
Fig. 5-3b: MFB (low tumble)
1
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0
-50 0 CAD 50 100
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0
-50 0 CAD 50 100
0% EGR 4% EGR 8% EGR 11% EGR
165
Fig. 5-3d: MFB (high tumble)
1
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0
-50 0 CAD 50 100
0% EGR 4% EGR 7% EGR
Figure 5-3 MFB verses CAD at KLMBT spark timings for a) the three intake airflow
conditions at 0% EGR, b) low tumble, c) swirl, d) high tumble
Gravimetric ISFCnet increased significantly for the swirl and high tumble conditions, as
shown in Figure 5-4; it was increased by 6.4 and 15.5%, respectively, as compared to the low
tumble condition. This is because the combustion duration was increased for the swirl
condition (Figure 5-3a), and the KLMBT spark timing retard caused the MFB50 (Figure 5-
3a) to be retarded from its optimum 8-10˚aTDC phasing for the high tumble condition (de O.
Carvalho et al., 2012). Thus, it is thought that these resulted in a reduced conversion
efficiency of the pressure and heat from the combustion into piston work, resulting in the fuel
consumption increases. The MFB50 retard for the high tumble condition was more
significant than the combustion duration increase for the swirl condition, explaining why its
fuel consumption was higher in comparison. Although the baffle plates for the swirl and high
tumble conditions did partially block the intake runner which alone would have increased
fuel consumption through increased pumping losses, the greater throttle angle required to
166
burn the additional injected fuel required for the swirl and high tumble conditions effectively
compensated for this; pumping work was only increased by 0.01 bar for the swirl condition
and it was decreased by 0.06 bar for the high tumble condition.
The fuel consumption generally decreased across the EGR range by 2.4, 3.9 and 10.2% for
the three engine conditions of low tumble, swirl and high tumble respectively. This is because
the KLMBT spark timings could be advanced towards their optimum phasing with the EGR
addition (de O. Carvalho et al., 2012). PMEP reductions (from 0.36 to 0.28 bar for the low
tumble condition, 0.37 to 0.30 bar for the swirl condition, and 0.29 to 0.27 bar for the high
tumble condition), and the reduced combustion temperatures which improved thermal
efficiency (Ratnak et al., 2015, Siokos et al., 2015) also contributed to the fuel consumption
improvements. All of these were explained previously in the first part of Chapter 4. The VAF
had an average error of 1.6%, as discussed in the previous chapter; therefore, the fuel
The fuel consumption increased slightly with 11% EGR addition at the low tumble condition
because the rate of KLMBT spark timing advance was reduced, when compared with the
advances achieved with other increases in EGR addition. The KLMBT spark timing was only
advanced by 2 CAD compared to 4 CAD with previous EGR additions, so the MFB50
became retarded as the KLMBT spark advance achieved was not sufficient to overcome the
laminar flame speed slowing effects of the additional EGR (Rhodes and Keck, 2012). This
caused its MFB50 to become retarded away from the optimum 8-10˚aTDC phasing (Figure 5-
167
Again, KLMBT spark timing advances would likely have been greater if intake manifold
temperature increases, resulting from the EGR addition, could have been eliminated; thus, the
300
290
280
ISFCnet (g/kWh)
270
260
250
240
0 2 4 6 8 10 12
EGR (%)
Low Tumble Swirl High Tumble
Figure 5-4 Gravimetric ISFCnet versus EGR ratio at KLMBT spark timings
5.3.5 EGT
From Figure 5-5 it can be seen that the EGT increased significantly by 16.0 and 68.2˚C at 0%
EGR addition, respectively, for the swirl and high tumble conditions, as compared to the low
tumble condition. It is proposed that the longer combustion duration for the swirl condition at
0% EGR addition (~24.0 CAD) as compared to the low tumble condition (~22.5 CAD),
resulted in the pressure and heat from the combustion being less efficiently converted into
piston work. This caused the gas temperature to become higher in the expansion and exhaust
strokes, resulting in the EGT increase observed. It is proposed that the significantly later
MFB50 for the high tumble condition at 0% EGR resulted in the same effect, producing the
observed behaviour. The greater MFB50 retardation for the high tumble condition as
168
compared to the other two intake airflow conditions (Figure 5-3a) was more significant than
the increased combustion duration for the swirl condition as compared to the low tumble
As to be expected, the EGT decreased with EGR addition for all of the intake airflow
conditions, because the MFB50 was more advanced (Figures 5-3b, 5-3c and 5-3d), as
Again, KLMBT spark timing advances would likely have been greater if intake manifold
temperature increases, resulting from the EGR addition, could have been eliminated. This
would likely have resulted in greater improvements in EGTs across the EGR range, as
740
720
700
EGT (°C)
680
660
640
620
600
0 2 4 6 8 10 12
EGR (%)
Low Tumble Swirl High Tumble
169
5.3.6 PM Emissions
The nucleation mode particles, identified as the smaller diameter particles forming the peak
on the left-hand side of the plot (10-30 nm) shown in Figure 5-6a, decreased significantly
with the swirl and high tumble conditions, as compared to the low tumble condition. The soot
accumulation particles, identified as the larger diameter particles forming the peak on the
right-hand side of the plot (30-300 nm), decreased too for the swirl condition, while they
remained approximately the same for the high tumble condition, as compared to the low
tumble condition. These significant changes have been confirmed by analysing the 95%
confidence intervals, which could not be displayed in the figures because the figures
It is proposed that the significant reduction in nucleation and accumulation particles observed
with the swirl condition was a result of not just the improved air-fuel mixture preparation due
to the KLMBT spark timing retard required (Table 5-2), but more importantly the increased
turbulence resulting from the swirl airflow. This provided more time for mixture preparation
as compared to the low tumble condition and it increased the rate at which the fuel and air
mixed, respectively. It is thought that this reduced the number of pockets of a high local
equivalence ratio meaning that the particulates oxidized more fully in the combustion,
resulting in the behaviour observed. The calculated average in-cylinder temperature profiles
for the swirl condition as compared to the low tumble condition (Figure 5-2a), at 0% EGR
addition, were similar with similar peaks; therefore, combustion temperature is not thought to
170
It is proposed that the significantly higher calculated average in-cylinder temperature for the
high tumble condition at 0% EGR addition (Figure 5-2a) resulted in the substantial reduction
of nucleation mode particles observed, along with the increased time available for air-fuel
mixture preparation resulting from the spark retard and the improved mixture formation
resulting from the high tumble airflow. However, the later MFB combustion phasing (Figure
5-3a) is thought to have reduced the time available for post-combustion oxidization, resulting
in similar numbers of accumulation mode particles as compared to the low tumble condition.
Overall, swirl had the greatest benefit regarding PN emissions, which means it has potential
regulations.
1.E+05
1.E+04
1.E+03
10 100 1000
Particle Diameter (nm)
Low Tumble Swirl High Tumble
171
Fig. 5-6b: PM (low tumble)
1.E+06
dN/dlogDp (#/cm3)
1.E+05
1.E+04
1.E+03
10 100 1000
Particle Diameter (nm)
0% EGR 4% EGR 8% EGR 11% EGR
1.E+05
1.E+04
1.E+03
10 100 1000
Particle Diameter (nm)
0% EGR 4% EGR 8% EGR 11% EGR
172
Fig. 5-6d: PM (high tumble)
1.E+06
dN/dlogDp (#/cm3)
1.E+05
1.E+04
1.E+03
10 100 1000
Particle Diameter (nm)
0% EGR 4% EGR 7% EGR
Figure 5-6 PN emissions versus particle diameter at KLMBT spark timings for a) the
three intake airflow conditions at 0% EGR, b) low tumble, c) swirl, d) high
tumble
The PN emissions for the low tumble, swirl and high tumble intake airflow conditions across
the EGR range are shown in Figures 5-6b, 5-6c and 5-6d, respectively. For the low tumble
condition, there was a significant decrease of nucleation and accumulation mode particles
across the EGR range. The significance of this and other PN changes discussed have been
confirmed by analysing the 95% confidence intervals, which could not be displayed in the
figures because the figures themselves would have become unclear. It is proposed that the
that this was more significant than the reduced time available for fuel-air mixing and thus
poorer mixture preparation with the KLMBT spark advances. The peak calculated average in-
cylinder temperature across the EGR range when comparing 0 and 11% EGR addition
(Figure 5-2b) remained approximately the same so it is not thought that this had an effect.
173
For the swirl condition, the PN did not change significantly across the EGR range, although
the accumulation mode particles did significantly increase as the EGR addition was increased
from 4 to 8%. It is proposed that the reduced calculated average in-cylinder temperatures
across the EGR range (Figure 5-2c) and the poorer air-fuel mixture preparation due to the
KLMBT spark timing advances (Table 5-2) across the range cancelled out the effect of the
increased post-combustion oxidization time and the effect of decreased primary carbon
particle formation with the reduced in-cylinder temperatures (Zhang et al., 2014b). This
caused the PN to remain approximately the same across the range, when comparing those
The PM for the high tumble intake airflow condition did not change significantly across the
EGR range. Despite the KLMBT spark timing being advanced as the EGR addition was
increased (Table 5-2), providing less time for fuel-air mixing, it is proposed that the improved
post-combustion oxidization time was sufficient to maintain the particles at similar levels
For the swirl condition, as mentioned, the accumulation mode particles increased
significantly as the EGR addition was increased from 4 to 8%. It is proposed that this was
caused by the increased calculated average in-cylinder temperatures (Figure 5-2c) which
increased primary carbon particle formation by the thermal pyrolysis and dehydrogenation
reaction of fuel vapour/droplets (Zhang et al., 2014b). As the EGR addition was further
increased to 11%, the calculated average in-cylinder temperature decreased (Figure 5-2c),
which resulted in the PM reducing due to the decreased primary particle formation.
If the intake manifold temperature increases with EGR addition could have been eliminated,
then it is expected that the PM emissions would have been reduced for the low tumble
174
condition due to an increased post-combustion oxidization time resulting from the expected
KLMBT spark timing improvements (Daniel et al., 2012e). However, the reduced fuel-air
mixing time could cancel out some of these improvements. This behaviour would also be
expected with the swirl and high tumble conditions, because similar behaviour was observed
in the experimental data when the KLMBT spark timing was advanced.
Overall, these results agree with the research of (Alger et al., 2010) and (Hedge et al., 2011)
who observed PN reductions with EGR addition, although it must be noted that their results
were both from a turbocharged engine; the varying enrichment of the fuel-air charge will
5.5 Conclusions
The effect of swirl and tumble intake airflows and hot EGR on the combustion and PM
emissions of the single-cylinder DISI research engine was investigated in this chapter and the
1. Swirl and high tumble significantly increased the fuel consumption with respect to the
low tumble condition by 6.4 and 15.5% respectively, but EGR addition minimized these
increases. The fuel consumption increase was 4.8 and 6.2%, respectively, at the
maximum EGR additions for swirl and high tumble conditions, compared to that
observed at the maximum EGR addition for the low tumble condition.
2. At 0% EGR addition, swirl significantly reduced the nucleation and accumulation mode
particles, with the high tumble intake airflow condition only significantly reducing the
175
nucleation mode particles. It is thought that the KLMBT spark timing retard required with
the swirl and high tumble intake airflow conditions increased the mixture preparation
time, resulting in the observed behaviour. Also, the additional turbulence produced by the
airflows improved mixture formation, reducing the number of pockets of a high local
oxidization time prevented the high tumble condition from reducing the accumulation
mode particles.
3. The PN emission was decreased significantly with EGR addition for the low tumble
significant change for the swirl and high tumble conditions. This is mainly because poorer
fuel-air mixture preparation caused by KLMBT spark timing advances cancelled out the
The main findings from this chapter are also summarized in Table 5-3.
Table 5-3 Results summary (intake airflow & EGR addition) (highlighted=improvement,
underlined=worsening)
Swirl [max. EGR] 6 CAD 3.9% 23.7˚C No sig. change No sig. change
(compared to 0% EGR)
High tumble [max. EGR] 8 CAD 10.2% 71.8˚C No sig. change No sig. change
(compared to 0% EGR)
176
Chapter 6
The aim of this chapter is to provide details of the combustion and emissions characteristics
of a DISI engine operated at different compression ratios with different fuels (8.5 bar IMEP),
and at different compression ratios with EGR addition (7.0 bar IMEP).
6.1 Introduction
The main combustion parameters investigated in this chapter are KLMBT spark timing, in-
cylinder pressure, calculated average in-cylinder temperature, MFB, fuel consumption and
exhaust temperature. PM, NOx and HC are the main emission parameters investigated.
There is an ever growing demand for reduction of PM, NO x and HC from engines as
emissions regulations become stricter, as well as the growing demand to reduce net CO2
emissions from the transportation sector, which is at the forefront of public perception. One
way to achieve these demands is to increase compression ratio; this improves thermal
efficiency causing fuel consumption and thus CO 2 output to reduce. Another way is to use
emissions. In addition, biomass is converted to make the renewable oxygenated fuel, which
reduces transportation’s net CO2 output (Wang et al., 2013). Furthermore, the Euro VI
emissions regulations which for the first time limit the PN from DISI engines have increased
interest in the effect of oxygenated fuels on engine particulates. They have the potential to
significantly reduce particulate emissions having health benefits, particularly for people
177
living in urban areas (Anderson et al., 2012b, United States Environmental Protection
Agency, 2014). The regulations have also increased interest in the effect of EGR addition on
engine particulates, with the potential of EGR coming under question because of it has been
observed to have a negative effect on them (Gill et al., 2011, Ma et al., 2014).
Overall, despite the amount of research that has been conducted into 1-butanol-gasoline and
ethanol-gasoline blended fuels, there appears to be lack of agreement in terms of the effect
these fuel blends on the combustion and emissions of gasoline engines. In addition, little
work has been conducted regarding the effect of these fuel blends on the combustion and
PM emissions. This is because the majority of the research conducted has been on PFI
fuel blends have not been studied well with each other along with a reference of gasoline fuel
at different compression ratios. Therefore, the research in this chapter has been conducted to
In addition, despite the amount of work that has been conducted into compression ratio and
EGR separately, little work has been conducted studying their combined effect on engine
these two fuel consumption (and NOx in the case of EGR) improving techniques on engine
particulates because of the limitations that the Euro VI emissions regulation has on these.
Therefore, these areas have also been investigated further in this chapter.
178
6.2 Experimental Procedure
The experimental test procedure outlined in Section 3.9 of this thesis was followed in order to
conduct the engine tests on the single-cylinder engine to collect the data presented and
discussed in this chapter. Relative air-fuel ratio λ was maintained at 1 during the experiments
and a COVIMEP of 5% was not exceeded. Valve timings were set at IVO=16˚bTDC and
EVC=36˚aTDC. KLMBT spark timings were used and the geometric compression ratio was
varied between 10.7 and 11.5 in the experiments. To change the compression ratio in the
engine the procedure outlined in Section 3.2.8 was followed. When using the non-gasoline
fuel splash blends the procedure discussed in Section 3.2.4.1 was followed to prevent
contamination between the fuels. In order to achieve hot EGR, and in order to achieve the
desired engine load and EGR addition, the equipment setup and experimental test procedure
The load of 8.5 bar IMEP was chosen to study for the investigation of the compression ratio
and fuel effect on engine combustion and emissions because it represents one of the worst
conditions for knock in this engine, as well as being an engine load that is highly relevant for
both NA and turbocharged DISI engines, increasing the usefulness of the data produced. The
load of 7.0 bar IMEP was chosen to study for the investigation of the compression ratio and
load in this engine where knock is present, and it allowed a reasonable, while not a high EGR
addition of 14% to be achieved. A higher engine load would mean that the throttle would
have needed to be more opened thus reducing the maximum level of EGR that could have
been achieved. One load was chosen to study for both parts of the investigation so that an in-
depth understanding could be formed; this would not be possible if multiple loads were
studied.
179
The spark was swept for all compression ratios, fuels and EGR additions in order to find the
KLMBT spark timing using the LabView on-line pressure trace analysis technique outlined
in Section 3.6.9. This allowed the KLMBT spark timing to be found much more quickly than
the MatLab post-experiment processing method. This is because it enabled the KLMBT spark
timings to be found while the engine was operating, thus allowing the measurements at the
KLMBT spark timings to be recorded alone, saving significant amounts of time. The SMPS
3936 was used to measure the PM emission for both parts of this chapter and the Horiba
MEXA-7100DEGR was used to measure the NOx and HC gaseous emissions for the first part
of this chapter. Gaseous emissions were not measured in the second part of this chapter
because it is believed that the literature sufficiently covers this area, and the EGT data was
A research grade gasoline was used in this research. This was used in addition to 1-butanol
and ethanol for the results in the first part of this chapter. All of the fuel properties as well as
those of the two tested fuel blends of Bu20 and E20 are outlined in Section 3.8. The ULG95
was used in its supplied form, while the 1-butanol and ethanol fuels were mixed with the
ULG95 fuel to form the Bu20 and E20 fuel splash blends with each containing 20%vol 1-
butanol and 20%vol ethanol respectively. The ULG95 fuel was supplied with 5%vol ethanol
pre-mixed in it, so the 20%vol 1-butanol blend and ULG95 fuel also had 5%vol ethanol in
them too, while the E20 blend had no ethanol in addition to the 20%vol. In total the Bu20
blend had 20%vol 1-butanol with 5%vol ethanol, the E20 blend had 20%vol ethanol and the
The test matrix provided in Table 6-1 was carried out in order to investigate the compression
ratio and fuel effect on engine combustion and emissions at the single engine load of 8.5 bar
IMEP. The results and discussion of these tests can also be found in (Lattimore, 2016b).
180
Table 6-1 Experiment test matrix (compression ratio & fuel)
Fuel
Bu20 1 2 3 4
E20 5 6 7 8
ULG95 9 10 11 12
For the investigation of the compression ratio and EGR effect on engine combustion and PM
emissions at the single engine load of 7.0 bar IMEP, the test matrix provided in Table 6-2
Table 6-2 Experiment test matrix (compression ratio & EGR addition)
From the KLMBT spark timings shown in Table 6-3, it can be seen that in the case of
gasoline, an increase in the compression ratio had no significant effect on KLMBT. The same
trend was also observed for the butanol blend (similar octane rating as gasoline) and even for
the ethanol blend, despite the high octane rating of ethanol. This is because at the engine load
of 8.5 bar IMEP, the engine was very prone to knock, even in the case of alcohols, due to the
high low temperature reactivity of alcohols (He et al., 2015) and the higher amount of fuel
181
being injected into the combustion chamber (i.e. ethanol has a lower calorific value than
butanol and gasoline). Thus, despite the compression ratio changing, no change in the
KLMBT spark timing could be realized. It must be noted that knocking could have been
avoided if the compression ratio was fully optimized for the fuels used. However, this was
not pursued in order to investigate the effect of the parameter changes on engine knock limit
It can also be seen that more advanced KLMBT spark timings could be achieved with Bu20
and E20 as compared to ULG95, with the most advanced spark timings being achieved with
Bu20. This is due to their higher octane numbers and the superior charge cooling effect of
alcohols compared to gasoline. Despite ethanol having a higher octane number than 1-butanol
and an increased cooling effect (in terms of mass), more advanced KLMBT spark timings
could be achieved with Bu20. It is believed that the 5%vol ethanol content in the Bu20 blend
(20%vol 1-butanol with 5%vol ethanol) was sufficient to compensate for the reduced charge
cooling effect and for the lower octane number of 1-butanol as compared to ethanol. It is also
thought that the higher chemical reactivity (He et al., 2015), faster laminar flame speeds
(Figure 6-5) and shorter fuel injection duration (less fuel quantity is required for the same
engine output power due to the higher heating value compared to ethanol) for the 1-butanol
blend with respect to the ethanol blend meant that the end-zone auto-ignition sites were
consumed before they had an opportunity to auto-ignite. Thus, these factors contributed to the
Overall, it is believed that E20 would perform better over a larger compression ratio range
than Bu20; particularly at compression ratios higher than 11.5, due to its higher octane
number and increased cooling effect. Thus, the small compression ratio range is a limitation
of this experiment.
182
Table 6-3 KLMBT spark timings (˚bTDC) (compression ratio & fuel)
Compression
Ratio 10.7 10.9 11.2 11.5
Fuel
Bu20 14 14 14 14
E20 12 12 12 12
ULG95 10 10 10 10
The in-cylinder pressure traces for the two fuels blends of Bu20 and E20 along with that for
the ULG95 reference fuel are shown in Figures 6-1a, 6-1b and 6-1c, respectively. It is clear
that as the compression ratio was increased, the maximum in-cylinder pressure increased, for
the two fuel blends and the reference fuel tested. This is because the more compact
combustion chamber achieved through the compression ratio increase, reduced the heat losses
to the surroundings, resulting in the in-cylinder pressure increases. The in-cylinder pressures
were highest for Bu20, followed by E20, then ULG95. This is due to the more advanced
KLMBT spark timings (Table 6-3) which could be achieved with Bu20 and E20 as compared
to those achieved with ULG95, with the most advanced spark timings being achieved for
Bu20. This resulted in heat being released from the fuel at an earlier CAD, enabling higher
in-cylinder pressures to be achieved before the downwards piston motion caused the in-
183
Fig. 6-1a: In-cylinder Pressure (Bu20)
45
35
30
25
20
15
10
0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
40
35
30
25
20
15
10
0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
184
Fig. 6-1c: In-cylinder Pressure (ULG95)
45
35
30
25
20
15
10
0
-50 0 CAD 50 100
Figure 6-1 In-cylinder pressure versus CAD for a) Bu20, b) E20 and c) ULG95 at
KLMBT spark timings
Figures 6-2a, 6-2b and 6-2c show the calculated average in-cylinder temperatures for the two
fuel blends of Bu20 and E20 and for the reference fuel of ULG95, respectively. Overall, the
calculated average in-cylinder temperature increased as the compression ratio was increased.
This is because of the aforementioned more compact combustion chamber achieved with the
compression ratio increase reducing heat losses to the surroundings. The calculated average
in-cylinder temperatures were highest for ULG95, with Bu20 and E20 having lower but
similar calculated average in-cylinder temperatures across the compression ratio range. It is
proposed that this is due to the higher heat of vaporization of 1-butanol and ethanol as
compared to ULG95 (Table 3-9). This meant that more energy was required to vaporize these
fuels, causing the average in-cylinder temperatures to reduce. The reduced fuel calorific
values of Bu20 and E20 compared to ULG95 also contributed to the lower average in-
185
Fig. 6-2a: In-cylinder Temperature (Bu20)
2100
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
1900
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
186
Fig. 6-2c: In-cylinder Temperature (ULG95)
2100
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
Figure 6-2 Calculated average in-cylinder temperature versus CAD at KLMBT spark
timings for a) Bu20, b) E20 and c) ULG95
6.3.1.3 MFB
The MFB profiles for the two tested fuel blends of Bu20 and E20, and the reference fuel of
ULG95 are shown in Figures 6-3a, 6-3b and 6-3c, respectively. For E20, there are no
significant differences between the profiles at the different compression ratios while Bu20
and ULG95 show a slightly advanced combustion as the compression ratio was increased. It
is proposed that the more highly compressed fuel-air mixtures at the higher compression ratio
burned more quickly than the less highly compressed mixtures at the lower compression
ratios, causing the combustion to proceed more quickly. Despite this, it appears that the last
stage of combustion (less than 10% of the fuel mass remaining) was faster at lower
compression ratios for all three fuels. For a quantitative analysis of the combustion speed, the
MFB50 and MFB10-90 have been calculated from the MFB profiles (please see next
section).
187
Fig. 6-3a: MFB (Bu20)
1.0
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
188
Fig. 6-3c: MFB (ULG95)
1.0
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
Figure 6-3 MFB versus CAD at KLMBT spark timings for a) Bu20, b) E20 and c)
ULG95
Figure 6-4 shows the MFB50 data for the two tested fuel blends of Bu20 and E20, and the
tested reference fuel of ULG95, across the compression ratio range. As discussed and
explained previously, the KLMBT spark timings were most advanced for Bu20, with E20
second and ULG95 third, thus leading to the most advanced MFB50 of Bu20 across the
compression ratio range, followed by E20 and ULG95. The MFB50 remained almost
constant across the compression ratio range for E20; this is reflected in the MFB profile for
E20 (Figure 6-3b). It is thought that experimental error resulted in the observed trend, as
indicated by the confidence intervals. However, for the other two fuels of Bu20 and ULG95,
there was a significant reduction in the MFB50 across the compression ratio range. Again, it
is proposed that the more highly compressed fuel-air mixtures at the higher compression ratio
burned more quickly than the less highly compressed mixtures at the lower compression
ratios, causing the combustion to proceed more quickly, thus resulting in the MFB50
becoming advanced.
189
23
22
21
MFB50 (˚aTDC)
20
19
18
17
16
15
10.6 10.8 11.0 11.2 11.4 11.6
Compression Ratio
Bu20 E20 ULG95
Figure 6-5 shows the combustion duration (MFB10-90) data for the two tested fuel blends of
Bu20 and E20, and the tested reference fuel of ULG95, across the compression ratio range. 1-
butanol and ethanol addition to gasoline reduced the combustion duration of the fuel; it is
proposed that 1-butanol and ethanol increased the laminar flame speed, due to the oxygen in
their molecules. The higher chemical reactivity of 1-butanol as compared to ethanol and the
shorter injection duration of Bu20 with respect to E20 explain its shorter combustion duration
in comparison. The shorter combustion duration of Bu20 and E20 as compared to ULG95 is
also due to enhanced in-cylinder turbulence (Stone, 1999) during their combustion which
results from their advanced spark timings. Furthermore, as a result of their advanced spark
timings, the combustion occurs in a smaller volume which increases the in-cylinder
temperature that further promotes combustion. It has to be also noted that the combustion
duration of Bu20 reduced significantly across the compression ratio range; this continues the
trend in Figure 6-4 which shows that the first half of the combustion process also proceeded
190
If the compression ratio was fully optimised for each fuel tested, the combustion phasing
would have been optimised for each fuel. Therefore, it would be expected that the
combustion of each fuel, particularly that of ULG95, would have been quicker due to higher
24.5
23.5
MFB10-90 (CAD)
22.5
21.5
20.5
19.5
18.5
10.6 10.8 11.0 11.2 11.4 11.6
Compression Ratio
Bu20 E20 ULG95
The indicated efficiency for the two tested fuel blends of Bu20 and E20, and the tested
reference fuel of ULG95, across the compression ratio range, is shown in Figure 6-6. They
increased by 1.3, 1.3 and 1.1% for Bu20, E20 and ULG95, respectively. This compares to a
maximum theoretical thermal efficiency increase of 1.8% which can be obtained from
Equation 6-1 by assuming γ=1.4 and solving for the minimum and maximum respected
(6-1)
191
Therefore, the thermal efficiency increase observed is realistic. It is well known that the
volumetric fuel consumption of oxygenated fuels and their low percentage blends in gasoline
is significantly higher than that of gasoline due to their lower calorific value. Therefore, this
is why the indicated efficiency rather than the volumetric fuel consumption has been analysed
in this section.
producing the observed behaviour. Bu20 had the highest indicated efficiency, followed by
E20 then ULG95, due to their respected KLMBT spark timings (Table 6-3) and their
respected combustion durations (Figure 6-5). The more advanced the spark timing and the
faster the combustion, the more efficiently the fuel was converted into engine power, thus
resulting in the indicated efficiency increases observed. Their higher indicated efficiencies
also resulted from their lower combustion temperatures (Figures 6-2a, 6-2b and 6-2c) which
It should be noted that the compression ratio of this engine is not optimised. Within the
compression ratio range tested, the knock limit was reached before the optimal MFB50
phasing was achieved (Figure 6-4). If the fuels were tested across a range of lower
compression ratios, it is expected that knock may not occur even at WOT. Thus, at the tested
engine load of 8.5 bar IMEP, the engine efficiency may be improved for the fuels tested
because the combustion phasing will be optimized and there will be a higher in-cylinder
turbulence at the time of ignition. This would be expected to reduce the combustion duration
and thus improve thermal efficiency further. However, it is important to note that there is a
trade-off because thermal efficiency would be reduced at lower compression ratios. It must
also be noted that because the injected fuel mass is increased with the Bu20 and E20 fuel
blends as compared to ULG95, a fuel pump in a real engine would consume more power
192
which would reduce the indicated efficiency in comparison to that observed in this
investigation.
0.350
0.345
Indicated Efficiency
0.340
0.335
0.330
0.325
0.320
10.6 10.8 11.0 11.2 11.4 11.6
Compression Ratio
Bu20 E20 ULG95
Figure 6-6 Indicated efficiency versus compression ratio at KLMBT spark timings
6.3.1.6 EGT
Figure 6-7 shows the EGTs for the two tested fuel blends of Bu20 and E20, and the tested
reference fuel of ULG95 across the compression ratio range. It is clear to see that there is
general small decrease in EGTs across the compression ratio range. It is proposed that as the
compression ratio was increased and the MFB50 became advanced, the pressure and heat was
more efficiently converted into piston work leading to the EGT decreases across the
compression ratio range. (Kramer et al., 2000) also observed EGT reductions as compression
The results also show that ULG95 had the highest EGT for all compression ratios, followed
by E20, then Bu20. It is proposed that the more advanced MFB50 of Bu20 as compared to
ULG95 and E20 (Figure 6-4) resulted in more efficient conversion of the pressure and heat
into work on the piston, resulting in the reduced EGTs in comparison. The MFB50 phasing
193
was more advanced for E20 than ULG95 for all compression ratios (Figure 6-4) leading to
lower EGTs in comparison. The lower calculated average in-cylinder temperatures for the
Bu20 and E20 fuel blends due to their higher heat of vaporization as compared to ULG95,
695
690
685
680
675
EGT (˚C)
670
665
660
655
650
645
10.6 10.8 11.0 11.2 11.4 11.6
Compression Ratio
Bu20 E20 ULG95
6.3.1.7 PM Emissions
The PN emissions for the two tested fuel blends of Bu20 and E20, along with the tested
reference fuel of ULG95 are shown in Figures 6-8a, 6-8b and 6-8c, respectively. It is clear to
see from Figure 6-8a that the compression ratio increase reduced the smaller nucleation mode
particles on the left-hand side of the plot (10-30 nm) for the Bu20 blend. It is proposed the
observed reduction was due to the increased calculated average in-cylinder temperatures
across the compression ratio range which increased the oxidation of the particles in the
combustion chamber. The KLMBT spark timing was unchanged across the compression ratio
194
range, therefore, mixture preparation was not considered to have had an effect on the
observed behaviour. Although the in-cylinder temperature increases would have increased
primary carbon particle formation by the thermal pyrolysis and dehydrogenation reaction of
fuel vapour/droplets (Zhang et al., 2014b), the increased oxidation of the particles also
resulting from the in-cylinder temperature increases was clearly the stronger effect.
E20 showed a similar trend to Bu20 but it was much weaker; the nucleation mode particles
decreased as the compression ratio was increased. Again, it is proposed that the higher
calculated average in-cylinder temperatures (Figure 6-2b) increased the rate of oxidation of
these particles in the combustion chamber, leading to the observed trend. For both Bu20 and
E20, no significant changes in accumulation mode particle numbers on the right hand side of
the plot (30-300 nm) were observed. It is believed that the increased oxidization of particles
resulting from the increased calculated average in-cylinder temperatures across the
compression ratio range was cancelled out by increased rate of particle formation caused by
the increased primary carbon particle formation (Zhang et al., 2014b), also resulting from the
The data for ULG95 shows a completely uni-modal distribution with no significant
nucleation mode particles being recorded. As the compression ratio was increased, the
mode particles increased across the compression ration range for ULG95 because the
formation (Zhang et al., 2014b), as with E20. This appears to have overcome the effect of
increased particle oxidization resulting from the higher calculated average in-cylinder
temperatures. Again, the KLMBT spark timing was unchanged across the compression ratio
range, thus mixture preparation is not thought to have had an effect on the observations.
195
For the two tested fuel blends Bu20 and E20, along with the tested reference fuel ULG95, it
is proposed that significant nuclei adsorption of nucleation mode particles onto the
accumulation mode particles occurred and this along with the thermodenuder, which removed
many of the nucleation particles before they could be measured, led to the mostly uni-modal
behaviour observed.
Overall, the effect of compression ratio increase on PM emissions is not significant when the
95% confidence intervals are taken into consideration, due to the compression ratio increase
being relatively small. The 95% confidence intervals themselves could not be displayed on
1.E+05
1.E+04
10 100 1000
Particle Diameter (nm)
CR=10.7 CR=10.9 CR=11.2 CR=11.5
196
Fig. 6-8b: PM (E20)
1.E+06
dN/dlogDp (#/cm3)
1.E+05
1.E+04
10 100 1000
Particle Diameter (nm)
CR=10.7 CR=10.9 CR=11.2 CR=11.5
1.E+05
1.E+04
10 100 1000
Particle Diameter (nm)
Figure 6-8 PN emissions versus particle diameter at KLMBT spark timings for a) Bu20,
b) E20 and c) ULG95
197
6.3.1.7.2 Fuel Effect on PN Emission
Comparing the behaviours of the different fuels in Figures 6-8a, 6-8b and 6-8c, 1-butanol
significantly reduced the PN when added to the gasoline fuel, whereas ethanol had little or no
effect. It is proposed that the significantly earlier MFB50 phasing (Figure 6-4) and shorter
combustion duration of Bu20 (Figure 6-5), as compared to the other two fuels, provided more
time for oxidation of the particulates after the combustion process, leading to the significant
PN reduction. This appears to have overcome effect of the advanced KLMBT spark timing
(Table 6-3), which would have reduced the fuel-air mixing time, as well as the effect of the
reduced calculated average in-cylinder temperatures (Figure 6-2a). These would have
resulted in more pockets of a high local equivalence ratio and a reduced oxidation rate in the
Also, it is important to note that reduced calculated average in-cylinder temperatures will
have also reduced the soot formation rate through reducing the formation of primary carbon
particles (Zhang et al., 2014b), thus contributing to the reductions observed. In addition, it is
thought that because the gasoline already had 5%vol ethanol content, the increase in ethanol
content to 20%vol made little difference to the PN behaviour of E20. Furthermore, E20 may
not have reduced PM because of increased piston/wall wetting resulting from the increased
injection duration required to maintain the same engine load, due to its reduced calorific
value. The calorific value of Bu20 is higher so this effect was not as significant as with E20.
Overall, there is no significant effect of fuel type on the particles average size with all
198
(Gu et al., 2012, He et al., 2010, Karavalakis et al., 2013, Niass et al., 2012, Zhang et al.,
2014b) also reported that 1-butanol addition to gasoline fuel reduced the PN concentration,
and (Bielaczyc et al., 2014, Catapano et al., 2013, Catapano et al., 2014, Costagliola et al.,
2013, Di Iorio et al., 2011, Ojapah et al., 2014, Storey et al., 2010, Vuk and Vander Griend,
2013) also observed the same for ethanol addition to gasoline fuel.
There are further reasons as to why the accumulation mode particles decreased with 1-
butanol addition to gasoline fuel. There is a positive correlation between the accumulation
mode particles and the polycyclic aromatic hydrocarbons (PAHs); the addition of alcohol to
gasoline reduces the aromatic content of the fuel, thus it also caused the accumulation mode
particles to decrease (De Abrantes et al., 2009, Zhang et al., 2014b). Furthermore, the oxygen
content in the fuel blend leads to a lower formation rate of soot and also to a higher oxidation
rate of soot (Zhang et al., 2014b). Biofuels tend to have a lower droplet velocity and
relatively low mean droplet diameter, which reduces piston impingement resulting in higher
combustion efficiency (Figure 6-6) and lower accumulation mode particle formation (Tian et
al., 2010). Furthermore, biofuels burn at higher pressures (Figures 6-1a and 6-1b) which help
Lastly, Bu20 had a noticeably higher number of nucleation mode particles than the other two
fuels tested. It is thought that this was due to the lower soot accumulation mode particles
observed, which meant less adsorption of the nucleation mode particles onto the
accumulation mode particle surfaces occurred, leading to higher numbers being observed.
Overall, the effect of 1-butanol addition to gasoline on PM emissions is significant when the
95% confidence intervals are taken into consideration, while ethanol addition to gasoline has
199
6.3.1.8 Gaseous Emissions (NOx and HC)
Figure 6-9 presents the NOx emission data for the two tested fuel blends and tested reference
fuel. Overall, there is a significant increase in NOx emissions across the compression ratio
range; they increased by 17.9% for Bu20, 21.7% for E20 and 23.5% for ULG95. These
temperatures across the compression ratio range, which caused more NO x to be formed. It is
also clear that ULG95 had the highest NOx emission, followed by Bu20 then E20. It is
proposed that the lower calculated average combustion temperatures of Bu20 (Figure 6-2a)
and E20 (Figure 6-2b) reduced the formation of NOx emissions in comparison to ULG95.
Despite the calculated average in-cylinder temperatures being similar for Bu20 and E20
across the compression ratio change and ethanol having a higher oxygen to carbon ratio than
1-butanol, Bu20 produced more NOx emissions than E20. It is proposed that the earlier
MFB50 of Bu20 as compared to E20 (Figure 6-4) resulted in the observed behaviour. With an
advanced MFB50, the fuel-air mixture burns at an earlier point in the engine cycle where the
thought that this resulted in higher local temperatures which resulted in the increased NO x
emissions for Bu20. Furthermore, the NOx emissions can also be attributed to the H/C ratio.
E20 had the highest H/C ratio, followed by Bu20 then ULG95. As reported by (Harrington
and Shishu, 1973), the NOx emission has an inverse relationship to H/C. The combustion of
fuels with a higher H/C ratio tends to have a lower adiabatic flame temperature because water
has a higher specific heat capacity than CO2 (Harrington and Shishu, 1973, Daniel et al.,
2011), thus further explaining why the NOx emission formation was lower for E20 as
compared to Bu20.
200
11.0
10.5
10.0
9.5
NOx (g/kWh)
9.0
8.5
8.0
7.5
7.0
6.5
10.6 10.8 11.0 11.2 11.4 11.6
Compression Ratio
Bu20 E20 ULG95
Figure 6-9 NOx emissions versus compression ratio at KLMBT spark timings
Figure 6-10 presents the HC emissions data for the two tested fuel blends and the tested
reference fuel. It is clear to see that the HC emissions increased significantly across the
compression ratio range; they increased by 20.9% for Bu20, 20.8% for E20 and 26.2% for
ULG95. It is believed that the increased surface to volume ratio of the combustion chamber
and the higher relative influence of the crevice volume as compared to the whole volume of
the combustion chamber resulted in the observed HC emission increases (Kramer et al., 2000,
The HC emissions were lower for Bu20 and E20 as compared to ULG95 because their
oxygen content was higher, which promoted the oxidation of HC in the combustion chamber.
This appears to have overcome the reduced fuel-air mixing time caused by the more
advanced KLMBT spark timings and the reduced combustion temperatures, which alone
would have caused the HC emissions to increase. Ethanol has a higher oxygen to carbon ratio
than 1-butanol, thus, there was a higher HC oxidation rate with E20 as compared to Bu20,
leading to lower HC emissions in comparison. Also, the KLMBT spark timing was more
201
advanced for the Bu20 fuel blend in comparison to E20, resulting in poorer mixture
preparation and thus higher HC emissions. Despite the differences observed in the HC (and
NOx) emission between the different fuels, this would be expected to diminish in a vehicle
It must be noted that these HC emissions are uncorrected; the sensitivity of the FID analyser
to oxygenated compounds has not been taken into account. Research has shown that FID
analysers have reduced sensitivity to oxygenated fuels (Cheng et al., 1998, Wallner and
Miers, 2008), suggesting that the HC emissions for Bu20 and E20 are higher than those
observed in these results, necessitating detailed HC emission speciation for reliable analysis.
The ULG95 fuel also contains oxygenated components (5%vol ethanol content) so its HC
5.1
4.9
4.7
4.5
HC (g/kWh)
4.3
4.1
3.9
3.7
3.5
10.6 10.8 11.0 11.2 11.4 11.6
Compression Ratio
Bu20 E20 ULG95
202
6.3.1.9 Big Picture
Figure 6-11 shows the overall effect of compression ratio and fuel on the gaseous emissions,
indicated efficiency and total PN. It is clear to see that for ULG95, the gaseous emissions of
NOx and HC increased with increased compression ratio, along with the indicated efficiency
and total PN. However, when 1-butanol and ethanol are blended into the ULG95 fuel, the
gaseous emissions of NOx and HC are reduced, along with total PN, and the indicated
efficiency is increased. Ethanol is most effective to reduce the gaseous emissions (NOx and
HC) of the ULG95 fuel and 1-butanol is most effective to reduce the total PN emission.
Figure 6-11 Overall effect of compression ratio and fuel on gaseous emissions, indicated
efficiency and total PN (integrated across 10-289 nm range) at KLMBT spark
timings
203
6.3.2 Investigation of the Compression Ratio and Hot EGR Effect on
Combustion and PM Emissions
The KLMBT spark timings for the three EGR conditions are shown in Table 6-4. As the
compression ratio was increased, the KLMBT spark timing was retarded because the fuel-air
mixture was more compressed at TDC. This increased the pressure and temperature in the
combustion chamber, leading to increases in the end-zone temperatures, which would have
resulted in engine knock had the spark timing not been retarded. The research engine began
to knock at a load of 6.0 bar IMEP, thus, the compression ratio increase made the engine
more prone to knock at the tested load of 7.0 bar IMEP, resulting in the spark retard required.
The KLMBT spark timing was advanced as the EGR addition was increased because the
temperatures in the end-zone of the combustion chamber were suppressed by the EGR gases.
Despite this, KLMBT spark timing advances would likely have been greater if intake
manifold temperature increases, resulting from the EGR addition, could have been
eliminated. The overall effect of the compression ratio increase and EGR addition was to
advance the spark timing by 2 CAD, due to the knock suppression ability of the EGR gases.
The 4 CAD retard in KLMBT spark timing at 7% EGR addition as the compression ratio was
increased from 11.2 to 11.5, larger than the 2 CAD spark retard required with the other
204
Table 6-4 KLMBT spark timings (˚bTDC) (compression ratio & EGR addition)
The in-cylinder pressures at the KLMBT spark timings, as shown in Figures 6-12a, 6-12b and
6-12c for the three EGR conditions, generally reduced as the compression ratio was increased
and the KLMBT spark timings (Table 6-4) had to be consequently retarded to avoid engine
knock. It must be noted that knocking could have been avoided if the compression ratio was
fully optimized for the fuel used. However, this was not pursued in order to investigate the
effect of the parameter changes on engine knock limit and the consequent combustion and
At the compression ratio of 10.9, the in-cylinder pressure at 0 and 7% EGR addition did not
reduce as compared to that at the compression ratio of 10.7, and it increased at 14% EGR
addition. It is proposed that although the KLMBT spark timing had to be retarded with the
compression ratio increase, the rate of heat release (ROHR) increased due to the more highly
compressed fuel-air charge. At 14% EGR addition, this was also sufficient to overcome the
retarded KLMBT spark timing, as the compression ratio was increased to 10.9. However,
further increases in the ROHR produced by compression ratio increases beyond 10.9 were not
sufficient to overcome the retarded combustion phasing resulting from the KLMBT spark
timing retards, causing the in-cylinder pressure to reduce as the compression ratio was further
increased.
205
The overall effect of compression ratio increase and EGR addition was reduced in-cylinder
pressures as can be seen by comparing Figures 6-12a and 6-12c. It would appear the reduced
ROHR produced by the EGR addition was more significant than the advanced combustion
phasing resulting from the 2 CAD KLMBT spark timing advance (Table 6-4). Similarly
(Alger et al., 2015) observed that in-cylinder pressure increases resulting from an increased
compression ratio were cancelled out by the consequent spark retard required to avoid engine
knock.
Smaller in-cylinder pressure decreases with EGR addition would have been likely if intake
manifold temperature increases, resulting from the EGR addition, could have been
eliminated. This is because, as mentioned, it would likely have led to a greater KLMBT spark
timing advance and thus more of the energy released from the fuel would have contributed to
40
35
30
25
20
15
10
0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
206
Fig. 6-13b: In-cylinder Pressure (~7% EGR)
45
35
30
25
20
15
10
0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
40
35
30
25
20
15
10
0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
Figure 6-12 In-cylinder pressure versus CAD at KLMBT spark timings for a) 0% EGR, b)
7% EGR, c) 14% EGR
The calculated average in-cylinder temperatures at the KLMBT spark timings for the three
EGR conditions, as shown in Figures 6-13a, 6-13b and 6-13c, generally reduced with
compression ratio increases and consequent spark timing retards. This is again due to the
KLMBT spark timing retards required with the compression ratio increases, as with the in-
cylinder pressure reductions. The overall effect of increased compression ratio and EGR
207
addition was to reduce the average calculated in-cylinder temperatures as can be seen by
comparing Figures 6-13a and 6-13c. This is the result of the KLMBT spark timing retards
required to avoid engine knock at the increased compression ratios and the temperature
suppression ability of the EGR gases due to their higher heat capacity as compared to air.
It is believed that the increased intake manifold temperatures resulting from the EGR addition
were compensated for by the less optimized combustion phasing resulting from the
consequently reduced KLMBT spark timing advance, compared to the maximum spark
advance possible, as explained in the first part of Chapter 4. Thus, it is thought the two effects
of increased intake manifold temperature and reduced subsequent KLMBT spark timing
advances cancelled one another out, meaning there was no overall effect on the peak of the
calculated average in-cylinder temperatures. Again, it is thought that the improved KLMBT
spark timing would have quickly compensated for the intake gas temperature decrease by the
time temperatures reached the beginning of the range reported (500˚C); thus, the effect would
1900
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
208
Fig. 6-14b: In-cylinder Temperature (~7% EGR)
2100
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
1900
1700
1500
1300
1100
900
700
500
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
Figure 6-13 Calculated average in-cylinder temperatures versus CAD at KLMBT spark
timings for a) 0% EGR, b) 7% EGR and c) 14% EGR
6.3.2.3 MFB
The MFB profiles at the KLMBT spark timings for the three EGR conditions are shown in
Figures 6-14a, 6-14b and 6-14c. The profiles became retarded as the compression ratio was
increased because of the KLMBT spark retard that was required in order to avoid engine
knock. From the MFB profile for 14% EGR addition (Figure 6-14c) it can be seen that there
209
was a less significant difference between the profiles at different compression ratios as
compared to the MFB profiles for 0% and 7% EGR addition. It is thought that the knock
suppression ability of the EGR gases allowed the MFB50s at the different compression ratios
to be more advanced towards their optimum phasing, resulting in the MFB profiles becoming
From comparing Figures 6-14a and 6-14c, it can be seen that the overall effect of the
compression ratio increase and EGR addition was to retard and elongate the MFB profile, due
to the KLMBT spark timing retard required with the compression ratio increase and the
combustion slowing effects of the EGR gases (Rhodes and Keck, 1985).
Again, KLMBT spark timing advances would likely have been greater if intake manifold
temperature increases, resulting from the EGR addition, could have been eliminated; thus, the
retardation of the MFB profiles would likely have been reduced. The MFB profiles will have
become slightly more elongated too because the slightly cooler fuel-air charge will have
burned more slowly. If the compression ratio was fully optimised for the fuel used, the
combustion phasing would have been optimised for each EGR ratio tested, therefore, it would
be expected that the combustion would have been quicker at each condition due to higher in-
210
Fig. 6-15a: MFB (0% EGR)
1.0
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
211
Fig. 6-15c: MFB (~14% EGR)
1.0
0.9
0.8
0.7
0.6
MFB
0.5
0.4
0.3
0.2
0.1
0.0
-50 0 CAD 50 100
CR=10.7 CR=10.9 CR=11.2 CR=11.5
Figure 6-14 MFB versus CAD at KLMBT spark timings for a) 0% EGR, b) 7% EGR and
c) 14% EGR
Gravimetric ISFCnet at the KLMBT spark timings, as shown in Figure 6-15, increased across
the compression ratio range due to the KLMBT spark timing retards that were required. This
caused the MFB50 (Figures 6-14a-c) to become retarded away from its optimum 8-10˚aTDC
phase (de O. Carvalho et al., 2012), thus causing the fuel consumption increase observed.
Fuel consumption increased by 4.4, 4.5 and 4.4% for the 0, 7 and 14% EGR additions,
respectively. Therefore, increasing the compression ratio at 7.0 bar IMEP resulted in worse
Across the EGR range the fuel consumption decreased significantly, due to the KLMBT
spark timing advances and PMEP reductions that could be achieved. It decreased by 4.6, 4.6,
4.3 and 4.6% for the compression ratios of 10.7, 10.9, 11.2 and 11.5 respectively. While the
KLMBT spark timing advances were not sufficient to advance the MFB profile (Figures 6-
14a-c), the PMEP was reduced significantly across the EGR range. The PMEP reduced from
212
0.36 to 0.28 bar, 0.36 to 0.27 bar, 0.38 to 0.26 bar, and 0.34 to 0.25 bar for the compression
ratios of 10.7, 10.9, 11.2 and 11.5, respectively. Furthermore, the reduced combustion
temperatures with EGR addition improved thermal efficiency (Ratnak et al., 2015, Siokos et
Overall, the indicated fuel consumption decreased by 0.4% across the compression ratio and
EGR range; i.e. the fuel consumption was decreased by 0.4% at the compression ratio of 11.5
and EGR addition of 14% as compared to that observed at the compression ratio of 10.7 and
EGR addition of 0%. This is because of the reduction in pumping losses and the 2 CAD
KLMBT spark timing advance. Again, KLMBT spark timing advances would likely have
been greater if intake manifold temperature increases resulting from the EGR addition could
have been eliminated; thus, the fuel consumption reductions would likely have been greater.
noted that the VAF had an average error of 1.6%. Thus, the reduction of 0.4% could have
resulted from this error, meaning that the result observed is limited.
265
260
255
ISFCnet (g/kWh)
250
245
240
235
230
10.6 10.8 11.0 11.2 11.4 11.6
Compression Ratio
0% EGR 7% EGR 14% EGR
Figure 6-15 Gravimetric ISFCnet versus compression ratio at KLMBT spark timings
213
6.3.2.5 PM Emissions
The PN emissions at the KLMBT spark timings for the three tested EGR conditions are
shown in Figures 6-16a, 6-16b and 6-16c. For the 0 and 7% EGR additions, there was a
significant increase in nucleation mode particles across the compression ratio range. There
was also a small but significant increase in accumulation mode particles across the
compression ratio range at 0% EGR addition, while at 7% EGR addition, there was no
significant change in the accumulation mode particles across the range. Furthermore, the
accumulation mode particles significantly increased across the compression ratio range with
14% EGR addition, but there was no significant change in the nucleation mode particles at
It is proposed that the reduced calculated average in-cylinder temperatures (Figures 6-13a-c)
caused by the KLMBT spark timing retards (Table 6-4) reduced the oxidization of particles in
the combustion chamber in the expansion and exhaust strokes, resulting in the observed
behavior. This appears to have been more significant than the reduced primary carbon
particle formation, resulting from the reduced in-cylinder temperatures (Zhang et al., 2014b).
In addition, the time for post-combustion oxidization was reduced (Figures 6-14a-c) due to
the MFB profile becoming more retarded, contributing to the increased PNs observed. It is
believed that the accumulation mode particles did not significantly increase at 7% EGR
addition because the KLMBT spark timing became more retarded across the compression
ratio range (Table 6-4) as compared to the other EGR additions. This provided more time for
fuel-air mixing in comparison which reduced the number of pockets of a high local
equivalence ratio, meaning the fuel was burned more completely, cancelling out the effect of
214
Across the EGR range there was a small but significant decrease in the nucleation and
accumulation particles observed for the compression ratios tested. It is proposed that the
calculated average in-cylinder temperature decreases across the EGR range (Figures 6-13a-c)
reduced primary carbon particle formation by the thermal pyrolysis and dehydrogenation
reaction of fuel vapour/droplets, resulting in the PN decreases observed (Zhang et al., 2014b).
In addition, the slower combustion produced by the EGR addition (Figures 6-14a-c) provided
more time for the particulates to oxidize in the hot flames, contributing to the reductions
observed. These effects appear to have been more significant than the reduced fuel-air mixing
time due to the KLMBT spark timing advances (Table 6-4). This alone would have created
more pockets of a high local equivalence ratio which would have produced higher PNs.
The overall effect of the compression ratio increase and EGR addition was a small but
significant increase in the nucleation mode particle number and a significant reduction in the
accumulation mode PN. The KLMBT spark timing was only advanced by 2 CAD (Table 6-4)
therefore, it is not thought that the relative homogeneity of the two mixtures was significantly
different. However, the primary carbon particle formation was reduced (Zhang et al., 2014b),
due to the in-cylinder temperature decreases across the EGR range. This resulted in the
reduced PN observed. In addition, the slower combustion produced by the EGR addition
provided more time for the particulates to oxidize in the hot flames, contributing to the PN
reductions. It is thought that due to the lower numbers of accumulation mode particles across
the compression ratio range, less nuclei adsorption happened, which caused the observed
If the intake manifold temperature increases with EGR addition could have been eliminated,
then it is expected that the PM emissions would have been further reduced with EGR addition
due to the greater post-combustion oxidation time available with the expected greater
215
expected KLMBT spark timing advances (Daniel et al., 2012e). The slightly longer expected
combustion duration would also decrease the PM observed. However, poorer mixture
preparation resulting from the greater KLMBT spark timing advances would be expected to
limit the PM reductions, especially since this effect would be expected to result in an overall
PM increase in the results from the first part of Chapter 4. As discussed, the intake manifold
temperature effect on the calculated average in-cylinder temperature is believed to have been
1.E+06
1.E+05
1.E+04
10 100 1000
Particle Diameter (nm)
216
Fig. 6-17b: PM (~7% EGR)
1.E+07
dN/dlogDp (#/cm3)
1.E+06
1.E+05
1.E+04
10 100 1000
Particle Diameter (nm)
CR=10.7 CR=10.9 CR=11.2 CR=11.5
1.E+06
1.E+05
1.E+04
10 100 1000
Particle Diameter (nm)
Figure 6-16 PN emissions versus particle diameter at KLMBT spark timings for a) 0%
EGR, b) 7% EGR and c) 14% EGR
217
6.5 Conclusions
The effect of compression ratio and fuel on the combustion and emissions of a single cylinder
DISI research engine was investigated in this chapter and the following conclusions have
been made.
1. 1-butanol and ethanol addition to gasoline advanced the MFB50 phasing as well as
reducing the overall combustion duration across the compression ratio range; 1-
emission, mainly due to the earlier combustion phasing and thus increased post-
combustion oxidization time; ethanol addition to gasoline had little effect on the
emission.
emission across the compression ratio range, with ethanol being the most effective.
4. Overall, if combustion and PN emission parameters are the priority, then the Bu20
fuel blend has the most potential, while if NOx and HC emission parameters are the
priority, then the E20 fuel blend has the most potential. Synergies between
compression ratio increase and alcohol addition to gasoline enable PM and gaseous
The main findings from this section are also summarized in Table 6-5.
218
Table 6-5 Results summary (compression ratio & fuel) (highlighted=improvement,
underlined=worsening)
E20 (at compression ratio 0 CAD 1.3% 16.2˚C No sig. change No sig. change 21.7% 20.8%
of 11.5, compared to 10.7)
ULG95 (at compression 0 CAD 1.1% 13.4˚C No sig. change No sig. change 23.5% 26.2%
ratio of 11.5, compared to
10.7)
E20 (compared to ULG95 2 CAD 2.4% 14.6˚C No sig. change No sig. change 20.9% 12.1%
at compression ratio of
11.5)
The effect of compression ratio and hot EGR on the combustion and PM emissions of a
single cylinder DISI research engine was also investigated in this chapter and the following
5. EGR addition of 14% overcame the increased knock produced as the compression
ratio was increased from 10.7 to 11.5, resulting in an overall KLMBT spark timing
advance of 2 CAD with the compression ratio increases and EGR addition.
6. Fuel consumption was reduced significantly by 0.4% as the compression ratio was
increased from 10.7 to 11.5 and the EGR addition was increased from 0 to 14%;
efficiency) and the KLMBT spark timing advance resulted in the decrease observed.
7. PM was significantly reduced with compression ratio increase and EGR addition due
to the reduced primary carbon particle formation resulting from the in-cylinder
temperature decreases across the EGR range. In addition, the increased time available
219
for PM to oxidize in the hot flames, due to the slower combustion which resulted from
The main findings from this section are also summarized in Table 6-6.
220
Chapter 7
The main aim of this thesis was to study the effect of EGR on the combustion and emissions
of a DISI engine. In particular, the PM emissions were studied, and the results and discussion
from this work are shown in the preceding chapters. The simplicity of EGR should allow for
its adoption in all automotive gasoline engines rather than the partial adoption we see today,
allowing increased compression ratios to be used as well as increased swirl and tumble ratios.
However, PM still remains a challenge which could be overcome with increased swirl ratios
and by blending 1-butanol into gasoline fuel. The aim of this chapter is to present the
summary and conclusions of this current investigation and to make suggestions for future
work.
This thesis contains numerous results from three chapters of experimental data from a 4-
compression ratio of 10.7-11.5:1. The most significant findings of the investigation are
It must be noted that the compression ratio was not optimized for each fuel used and for each
engine condition tested. For the conditions which did not knock (e.g. 5.5 bar IMEP) the
compression ratio could have been increased to improve thermal efficiency and the optimum
combustion phasing would not have been affected. For the conditions which did knock, the
compression ratio could have been reduced in order to obtain the optimum combustion
phasing to improve thermal efficiency, although this would have been limited by the
221
reduction in thermal efficiency resulting from the compression ratio reduction. The least
optimized load was 8.5 bar IMEP when gasoline fuel was used with no EGR addition since
The Effect of EGR and its Type on Engine Combustion and Emissions
EGR addition of up to 13% enabled the KLMBT spark timing to be significantly advanced
due to the suppression of end-zone temperatures, which was sufficient to advance the MFB
profiles at 7.0 and 8.5 bar IMEP, but because the load of 5.5 bar IMEP was not knock
limited, the MFB profile was already optimized so improvements could not be made. The
more optimum MFB50 combustion phasing at 7.0 and 8.5 bar IMEP along with pumping
reductions at all tested loads due to the throttle being more opened reduced the fuel
consumption across the EGR ranges (2.2% at 5.5 bar IMEP, 4.1% at 7.0 bar IMEP, and 1.0%
at 8.5 bar IMEP). PM accumulation mode emissions generally increased at 5.5 and 7.0 bar
IMEP due to the lower in-cylinder and exhaust temperatures while they reduced in the
nucleation mode through increased nuclei adsorption due to the greater number of
accumulation mode particles present. The low EGR addition of 3% at 8.5 bar IMEP resulted
in no significant changes in the PM. Finally, NOx emissions decreased significantly due to the
lower in-cylinder temperatures and HC emissions increased significantly due to the lower in-
Extending the work further allowed the hot EGR system from the first experiment to be
compared against a cooled EGR system, along with a hot and cooled EGR system with a
simulated TWC, at 7.0 bar IMEP, with EGR addition of up to 14%. The KLMBT spark
timing could be most advanced with the cooled EGR after TWC EGR type due to the cooler
gases suppressing end-zone temperatures and because the EGR gases did not contain NO
which induces knock. The fuel consumption improvement (3.8%) was greatest for cooled
222
EGR, mostly due to its most improved MFB50 combustion phasing as compared to the other
EGR types. Gaseous emissions (NOx and HC) were best optimized with the cooled EGR after
TWC EGR type (45.3% NOx decrease and 10.1% HC increase). This is mainly because its
MFB50 combustion phasing did not advance across the EGR range as significantly as the
cooled EGR type. This resulted in less NOx formation in comparison, because with the less
advanced MFB50, the fuel-air mixture burned at a later point in the engine cycle where the
in-cylinder turbulence was weaker, causing the combustion to proceed less quickly. This is
believed to have resulted in lower local temperatures in comparison to the cooled EGR
condition which resulted in the lower NO x and thus greater NOx reductions across the EGR
range. Also, because the simulated gas came from a gas bottle, it did not introduce additional
HCs into the combustion chamber (reflecting the behaviour of a TWC with 100% efficiency)
as the hot and cooled EGR types did, reducing the HC increases observed.
The Effect of Intake Airflow and EGR on Engine Combustion and PM Emissions
Swirl and high tumble intake airflows significantly retarded the KLMBT spark timing at 7.0
bar IMEP. However, significant improvements could be made for all the intake airflow
conditions with EGR addition of up to 11%. Fuel consumption increased significantly with
swirl (6.4%) and high tumble (15.5%) intake airflows as compared to the reference low
tumble condition because of the KLMBT spark timing retards required. However, the
improvements made with EGR addition reduced the fuel consumption penalty significantly
compared to the reference condition (4.8% and 6.2% increase, respectively, at maximum
EGR additions).
PM was significantly reduced with the swirl intake airflow due to improved mixture
formation resulting from the KLMBT spark timing retard providing more fuel-air mixing
time and from the increased in-cylinder turbulence produced by the airflow. The high tumble
223
condition resulted in reduced nucleation but similar accumulation mode particles compared to
the reference low tumble condition. This is because the post-combustion oxidization time was
significantly reduced by the KLMBT spark timing retard and the improved mixture formation
was not sufficient to overcome this. EGR addition reduced PM for the reference low tumble
condition due to an increased post-combustion oxidization time resulting from the KLMBT
spark timing advances. However, it remained approximately the same across the EGR range
for the swirl and high tumble conditions. It is thought that the KLMBT spark timing advances
resulted in poorer fuel-air mixture preparation due to reduced mixture preparation time
available, which in addition to the reduced calculated average in-cylinder temperatures for
the swirl condition across the EGR range, cancelled out the effect of the increased post-
combustion oxidization time. It also cancelled out the effect of decreased primary carbon
particle formation resulting from the in-cylinder temperature reductions for the swirl
condition.
The Effect of Compression Ratio, Fuel and EGR on Engine Combustion and Emissions
At 8.5 bar IMEP, compression ratio increases from 10.7 to 11.5 did not result in KLMBT
spark timing changes because the load condition of 8.5 bar IMEP was already prone to
knock, so it did not worsen despite increasing the compression ratio. Bu20 had the most
advanced KLMBT spark timing, followed by E20, then ULG95, mainly due to the 5%vol
ethanol pre-mixed in it, giving it a 25%vol total alcohol content, as compared to only 20%vol
for E20. Indicated efficiency was most improved for Bu20 and E20 due to their greater
KLMBT spark timing advances as compared to ULG95. Compression ratio had little effect
on the PM emission, with the most significant effect being the Bu20 fuel blend which
reduced PN. This was mainly due to its most advanced spark timing and shorter combustion
duration significantly increasing the post-combustion oxidation time available. Increasing the
224
ethanol addition from 5%vol (ULG95) to 20%vol (E20) had no significant effect on the PM
emission.
At 7.0 bar IMEP, compression ratio increases from 10.7 to 11.5 resulted in the KLMBT spark
timing becoming more retarded. This is because the engine was less prone to knock at this
load, so small changes in the compression ratio made a significant difference to the knock
tendency. EGR addition of 14% however enabled this effect to be overcome, enabling the
KLMBT spark timing to be advanced overall by 2 CAD with compression ratio increases and
EGR addition. Fuel consumption was significantly improved by 0.4% with the compression
ratio increases and EGR addition due to the reduced pumping losses with EGR addition and
the KLMBT spark timing advances. Finally, PM emissions were reduced overall with the
compression ratio increases and EGR addition, because of the improved oxidation of
particulates in the slower flame, and the reduced primary carbon particle formation resulting
DISI engines, along with that of swirl and tumble intake airflows, compression ratio increase
and oxygenated fuels. However, before the approaches outlined in this thesis can be fully
the behaviour of these parameters and to tackle the main challenges that have been identified
in their use. The author has thus outlined some of the outstanding issues below to serve as
In general, the investigation should be extended to quantify the effects of EGR and its type,
swirl and tumble intake airflows, compression ratio and oxygenated fuels over the entire
225
engine load and speed range. Split-injection strategies, different fuel injection timings and
different valve timings/lifts should be studied to optimize the combustion and emissions.
Synergies between these different techniques (e.g. swirl & tumble and EGR along with
increased compression ratio) should be investigated once they have all been explored
individually on a deep level. Results should be obtained not only at KLMBT spark timings
but at a 10 CAD retard from the KLMBT too, because these are used in cold-start for catalyst
The compression ratio should be optimized for each fuel to ensure that the MFB50 does not
become too retarded from its optimum phasing. However, it should be noted that any
decrease in the compression ratio will lead to lower indicated efficiencies at the engine loads
that did not knock while potentially increasing the indicated efficiencies at the higher engine
loads which did knock. This will only happen if the benefits of increased burning speed due
to the increased in-cylinder turbulence during the combustion are greater than the reduced
This investigation should be done ideally in a modern prototype development engine with
turbocharging to increase the usefulness of the data for current engine development. It should
Transient tests should be conducted including NEDC and the upcoming RDE tests both in
engine and vehicle test rigs to understand how the different techniques affect real-world
emissions.
Tests should be conducted to investigate whether the use of the thermodenuder makes the
nucleation mode particles recorded more consistent or not, since it is designed to remove
226
them. Furthermore, the tests should be conducted on 3 consecutive days in varying order to
take day-to-day variability, resulting from changing ambient conditions and engine drift
(Beck et al., 2006) into account, since this can potentially affect the data; especially PM.
Specific further investigation for each particular research area is detailed in Appendix C1.
227
Appendix A1: Analysis of the Uncertainties in the Recorded
Data
The data collected in this investigation is accurate; however, there will always be a limit to
how accurate the data is. Therefore, the accuracy of the main components of the engine test
rig and emissions measurements systems have been collected from the relevant manufacturer
instruction manuals and they are discussed in the following section. This builds on the work
of a previous research student in the engine group who conducted the initial analysis (Turner,
2010).
Control Technologies was used. The encoder had an accuracy which is better than 0.01% and
the drive can maintain the dynamometer speed to an accuracy of 0.1%. The VVT system had
a steady state positional accuracy of as shown on the LabView control panel for
the system. The Kistler 6041A water cooled pressure transducer had a linearity error of less
than % of the full scale. Its short term drift, as measured by a reference sensor in the
engine, when it is operating at 1500 rpm and at an engine load of 9 bar IMEP, quantified as a
pressure and as a percentage, were bar and %, respectively. The Kistler 5011B
charge amplifier used to amplify the pressure signal had a linearity error of less than %
The VAF meter produced AFR readings which had an average error of 1.6% for the AFRs
used in this investigation. The inaccuracies resulted mainly from the limited resolution of the
inclined u-tube manometer (1 mm) and the limited pressure difference produced by a given
228
Regarding the PM emission measurement, the DMS500 manufactured by Cambustion Ltd.
had a particle count accuracy of . For the SMPS system, the condensing particle
counter (model 3775) manufactured by TSi had a particle count accuracy of % for
particle concentrations lower than 5x104 particles/cm3 and % for particle concentrations
larger than 5x104 but smaller than 1x107 particles/cm3. The gaseous emission measurements
obtained with the Horiba MEXA-7100DEGR were accurate to within % of the full scale.
Overall, this data highlights the limited accuracy of the equipment used in this investigation;
particularly that of the particulate emission measurements. This should be taken into
229
Appendix B1: Data Continuity (Chapter 4)
The results from the 7.0 bar IMEP condition in the first part of the chapter are directly
comparable with the results from the hot EGR condition in the second half of the chapter,
because they were recorded at the same engine condition (7.0 bar IMEP, hot EGR). As can be
expected, there are some differences between the data. 14% EGR addition was achieved with
the hot EGR condition while only 12% EGR addition was achieved at the 7.0 bar IMEP
condition because two EGR lines rather than one were used, which enabled the flow rate of
EGR gases to be increased. As a result, the KLMBT spark timing advances (Table 4-4 vs.
Table 4-3) were greater and the EGT reductions were also greater (Figure 4-16 vs. Figure 4-
5), which resulted from the increased EGR addition and the subsequently improved KLMBT
Despite this, the MFB50 was retarded across the EGR range for the hot EGR condition in the
second part of this chapter (Figure 4-12) while it was advanced across the EGR range at the
7.0 bar IMEP condition (Figure 4-3b). This was due to the flame speed reducing in the final
10% MFB as can be seen by comparing the profiles. It is believed that the more advanced
KLMBT spark timing resulted in poorer mixture preparation, producing the observed
difference. The fuel consumption improvements were greater in the first part of the chapter
(Figure 4-4) with the 7.0 bar IMEP condition than they were in the second part of the chapter
with the hot EGR condition (Figure 4-14), despite the more advanced spark timing. It is
thought that the 1.6% error in fuel consumption measurements discussed previously resulted
The HC emission increase was greater in the first part of the chapter with the 7.0 bar IMEP
condition (Figure 4-8 vs. Figure 4-20), despite the lower EGR addition that was achieved for
230
the hot EGR condition in the second part of the chapter. It is believed that the research fuel
used in the second part of the chapter had a higher oxygen content than the pump fuel used in
the first part of the chapter, helping to suppress the HC emission increases more effectively
through improved HC oxidation. Experimental error may also account for some of the
difference observed.
231
Appendix B2: Data Continuity (Chapter 5)
The results from the 7.0 bar IMEP condition in the first part of Chapter 4 are directly
comparable with the results from the low tumble condition in this chapter, because they were
recorded at the same engine condition (7.0 bar IMEP, low tumble intake airflow). As can be
11% EGR addition was achieved in this chapter even though 12% EGR addition was
achieved in the first part of Chapter 4. This was deliberately done in order to achieve the
same maximum EGR addition as that for the swirl intake airflow condition in this chapter, to
enable a direct comparison to be made. The KLMBT spark timings (Table 5-2) were more
advanced than they were in the results from the 7.0 bar IMEP condition in the first part of
Chapter 4 (Table 4-3), most likely because of experimental error. The in-cylinder pressures
were increased more significantly across the EGR range (Figure 5-1b) than they were in the
first part of Chapter 4 (Figure 4-1b), because of the greater spark advance achieved (2 CAD
greater with 1% less EGR), in comparison. This is also why the calculated average in-
cylinder temperatures (Figure 5-2b) could be increased across the EGR range, while only
maintained across the EGR range in the results from the first part of Chapter 4 (Figure 4-2b).
The fuel consumption improvement (Figure 5-4) was less than that achieved in the first part
of Chapter 4 (Figure 4-4) because of the 1.6% error in fuel consumption measurements. The
EGT reductions (Figure 5-5) were greater than those achieved in the first part of Chapter 4
(Figure 4-5) because of the improved combustion phasing with EGR addition, resulting from
the greater KLMBT spark timing advances achieved. This improved the efficiency of
pressure and heat transfer into piston work in comparison, resulting in the greater EGT
reductions.
232
PM was reduced with EGR addition (Figure 5-6b) while it was increased with EGR addition
in the first part of Chapter 4 (Figure 4-6b), because of the calculated average in-cylinder
temperature increases achieved with EGR addition, while they could only be maintained
across the EGR range with EGR addition in the first part of Chapter 4. This improved the
oxidation of PM across the EGR range, resulting in the observed differences. The PN was
significantly lower than that observed in the first part of Chapter 4 mainly because of the
effect of the thermodenuder which was used with the SMPS 3936, while for the DMS 500 it
was not used. Differences in the equipment design and slight differences in the measuring
principle (i.e. differential mobility sizer method vs. scanning mobility particle sizer method)
Indeed, (Abdul-Khalek et al., 1999) found that the number count of particles less than 50 nm
in diameter are heavily influenced by the dilution conditions such as dilution ratio,
temperature, relative humidity and residence time, which would vary slightly between the
SMPS 3936 and DMS500. While those particles larger than 50 nm are not heavily influenced
by the dilution conditions, it is believed they would still be influenced by them to a certain
degree, thus helping to explain the observed differences between the data.
233
Appendix B3: Data Continuity (Chapter 6)
The results from the 7.0 bar IMEP condition in the first part of Chapter 4 are directly
comparable with the results from the compression ratio of 11.5 in the second half of this
chapter, because they were recorded at the same engine condition (7.0 bar IMEP,
compression ratio=11.5). As can be expected, there are some differences between the data.
The KLMBT spark timing advance with EGR addition in the second part of this chapter
(Table 6-4) is less than that achieved in the first part of Chapter 4 (Table 4-3) even though the
EGR addition was increased (2% increase). It is believed that experimental error accounts for
some of the differences observed. Also, the EGR line insulation was improved for the
experiments in this chapter, which increased the intake manifold temperature, contributing to
the decreased spark advance achieved, in comparison. The in-cylinder pressure decreased in
the results from the second part of this chapter (Figures 6-12a-c), while it increased slightly in
the first part of Chapter 4 (Figure 4-1b), with EGR addition, due to the reduced KLMBT
spark timing advances, in comparison. The increased intake manifold temperatures also
decreased the calculated average in-cylinder temperature reductions achieved with EGR
addition in the second part of this chapter, as compared to those achieved in the first part of
Chapter 4.
The PM accumulation mode decreased with EGR addition in the second part of this chapter
(Figures 6-16a-c) while it increased with EGR addition in the first part of Chapter 4 (Figure
4-6b), because of the smaller calculated average in-cylinder temperature reductions achieved
which resulted in improved PM oxidation across the EGR range. The reduced KLMBT spark
timing advance with EGR addition compared to that achieved in the first part of Chapter 4
234
The PN was significantly lower than that observed in the first part of Chapter 4 mainly
because of the effect of the thermodenuder which was used with the SMPS, while for the
DMS500 it was not used. Differences in the equipment design and slight differences in the
measuring principle (i.e. differential mobility sizer method vs. scanning mobility particle
sizer method) are also likely to have contributed to the lower PN.
Again, (Abdul-Khalek et al., 1999) found that the number count of particles less than 50 nm
in diameter are heavily influenced by the dilution conditions such as dilution ratio,
temperature, relative humidity and residence time, which would vary slightly between the
SMPS and DMS500. While those particles larger than 50 nm are not heavily influenced by
the dilution conditions, it is believed they would still be influenced by them to a certain
degree, thus helping to explain the observed differences between the data.
235
Appendix C1: Detailed Specific Further Investigation for
Each Research Area
A real TWC should be used in the suggested extended investigation because the simulated
TWC does not introduce any additional HCs into the engine, which has an effect on the
results. While a real TWC would convert most of the HCs, it would not be 100% efficient,
thus it would introduce some HCs back into the engine, affecting the combustion and
emissions. However, the EGR ratio variations resulting from the varying conversion
efficiency of the TWC would have to be taken into consideration in the results. A real TWC
also increases the exhaust back-pressure resulting in increased pumping losses which is
The EGR addition should also be increased using an exhaust back pressure valve, for
example, in order to quantify the effect of higher EGR ratios on the engine combustion and
emissions. The trade-off between increased exhaust back-pressure resulting from the valve
and higher EGR ratios should be studied to investigate whether it would be beneficial to
increase exhaust back-pressure in engines for the purpose of achieving increased EGR ratios.
This trade-off should also be investigated to see whether it is beneficial to remove the NO
from the EGR gases for knock reduction, considering any improvements in fuel economy
from subsequent KLMBT spark timing advances may be cancelled out by the effects of the
increased back-pressure from the TWC. EGR should be studied at a range of temperatures to
study the effect EGR temperature has on engine combustion and emissions in more detail.
The PM emissions should also be studied fully with the different EGR types; limited
equipment availability prevented this from being done in this investigation. The effect of the
236
thermodenuder on the PM emission should be more deeply investigated as well as the
differences between the DMS 500 and SMPS 3936 PM measurement equipment. Work
should also be extended to optical investigations, particularly with Planar Laser Induced
Fluorescence (PLIF), to investigate further the effect EGR has on soot formation.
Different swirl and tumble ratios should be tested in the suggested extended investigation in
order to optimize the ratios to achieve improved engine combustion and PM emissions
behaviour. Particularly, they should be further investigated in the optical DISI engine, in
order to fully quantify the effect of swirl and tumble ratios on the initial flame propagation
and the last remaining 10% MFB, in order to optimize the combustion process. The swirl and
tumble ratios should be quantified in the optical engine at 60 CAD intervals during the intake,
compression and ignition strokes in order to fully quantify the effect of the intake airflow on
the mixture preparation and combustion. A comparison should be made between the
estimated and actual swirl and tumble ratios to quantify the accuracy of the estimation
The PM change resulting from the retarded spark timings required with the additional intake
airflows of swirl and tumble and the PM change resulting from the increased in-cylinder
effect of these parameters on PM individually. Further to this, the relationship between flame
speed and knock tendency should be investigated further to find the optimum flame speed for
minimum knock.
An exhaust back pressure valve should be used on the thermal engine to enable higher EGR
ratios to be achieved to study the effect of the laminar flame speed increases resulting from
237
the increased swirl and tumble ratios being fully suppressed by the EGR gases. Finally,
butanol and ethanol fuels should be studied with swirl and tumble intake airflows since they
have the potential to compensate for the increased engine knock resulting from the faster
Compression Ratio
The engine should be tested with a greater compression ratio range because some of the
effects observed, particularly regarding PM behaviour, were not significant due to the
relatively small range tested in this investigation. In addition, it is believed that Bu20
performed better than E20 in this investigation because they were only tested over a small
compression ratio range. E20 would be expected to perform better than Bu20 over a larger
compression ratio range because of its superior properties (higher octane number, higher
HoV), particularly at compression ratios higher than 11.5. The larger compression ratio range
should also be investigated in a turbocharged DISI engine along with EGR, to quantify the
improvement in compression ratio that can be achieved with the knock suppression effect of
the EGR gases. This is because compression ratios in turbocharged engines are reduced in
comparison to their equivalent NA engines (Su et al., 2014), in order to achieve a high IMEP.
Thus, EGR has the potential to enable the compression ratio to be maintained or increased
Oxygenated Fuels
1-butanol should be tested with and without 5%vol ethanol pre-mixed in the gasoline fuel to
quantify exactly what effect the pre-mixed ethanol has on the engine combustion and
emissions. In addition, 1-butanol and ethanol should be tested at different blend ratios in
gasoline, including blends with both 1-butanol and ethanol in gasoline, in order to find which
blends provide the best combustion and emissions characteristics. This is important because
238
rather than competing with ethanol, 1-butanol and ethanol could be blended together in
gasoline to provide potential improvements over using just one of these fuels blended in
gasoline. The work should also be extended to optical investigations; particularly that of
laminar flame speed measurement, to quantify the effect of fuel blending on this parameter.
This is because laminar flame speed significantly influences KLMBT which consequently
affects indicated efficiency, exhaust temperature, and PM and gaseous emissions. A suitable
FTIR detector should be used when testing the oxygenated fuels to ensure that the
239
List of References
ALGER, T., HALL, M. AND MATTHEWS, R. 2000. Effects of Swirl and Tumble on In-
Cylinder Fuel Distribution in a Central Injected DISI Engine. SAE Technical Paper 2000-01-
0533, doi:10.4271/2000-01-0533.
ALGER, T., CHAUVET, T. AND DIMITROVA, Z. 2009a. Synergies between High EGR
Operation and GDI Systems. SAE Int. J. Engines 1(1):101-114, doi:10.4271/2008-01-0134.
ALGER, T. AND MANGOLD, B. 2009. Dedicated EGR: A New Concept in High Efficiency
Engines. SAE Int. J. Engines 2(1):620-631, doi:10.4271/2009-01-0694.
ALGER, T., GINGRICH, J., KHALEK, I., AND MANGOLD, B. 2010. The Role of EGR in
PM Emissions from Gasoline Engines. SAE Int. J. Fuels Lubr. 3(1):85-98, doi:10.4271/2010-
01-0353.
ALGER, T., GUKELBERGER, R., GINGRICH, J. AND MANGOLD, B. 2015. The Impact
of Cooled EGR on Peak Cylinder Pressure in a Turbocharged, Spark Ignited Engine. SAE
Int. J. Engines 8(2):455-463, doi:10.4271/2015-01-0744.
240
ANDERSON, J., LEONE, T., SHELBY, M., WALLINGTON, T., BIZUB, J. J. AND
FOSTER, M. 2012a. Octane Numbers of Ethanol-Gasoline Blends: Measurements and Novel
Estimation Method from Molar Composition. SAE Technical Paper 2012-01-1274,
doi:10.4271/2012-01-1274.
BECK, C., STEVENSON, P. AND ZIMAN, P. 2006. The Impact of Gasoline Octane on Fuel
Economy in Modern Vehicles. SAE Technical Paper 2006-01-3407, doi:10.4271/2006-01-
3407.
BENNETT, S. 2010. Diesel Engines- Modern Diesel Technology. New York: Delmar.
BIELACZYC, P., SZCZOTKA, A. AND WOODBURN, J. 2014. The Impact of Fuel Ethanol
Content on Particulate Emissions from Light-Duty Vehicles Featuring Spark Ignition
Engines. SAE Int. J. Fuels Lubr. 7(1):224-235, doi:10.4271/2014-01-1463.
BOCK, B., BELL, A. AND FLOWEDAY, G. 2013. Investigation into the Influence of
Charge Cooling and Autoignition Chemistry on the Greater Knock Resistance of Ethanol
over Iso-octane. SAE Technical Paper 2013-01-2615, doi:10.4271/2013-01-2615.
241
BURTON, G., HOLMAN, J., LAZONBY, J., PILLING, G. AND WADDINGTON, D. 2000.
Chemical Storylines. Oxford: Heinemann.
CAIRNS, A., BLAXILL, H. AND IRLAM, G. 2006. Exhaust Gas Recirculation for
Improved Part and Full Load Fuel Economy in a Turbocharged Gasoline Engine. SAE
Technical Paper 2006-01-0047, doi:10.4271/2006-01-0047.
CAMBUSTION LTD. 2011. DMS500 Fast Particulate Spectrometer with Heated Sample
Line High Ratio Diluter User Manual Version 3.5 Software Version UI v4.11.
CASTAGNÉ, M., DUMAS, J., HENRIOT, S. AND LAFOSSAS, F. A. 2003. New Knock
Localization Methodology for SI Engines. SAE Technical Paper 2003-01-1118,
doi:10.4271/2003-01-1118.
CHECKEL, M. AND DALE, J. 1989. Pressure Trace Knock Measurement in a Current S.I.
Production Engine. SAE Technical Paper 890243, doi:10.4271/890243.
CHEN, R., OKAZUMI, R., NISHIDA, K. AND OGATA, Y. 2015. Effect of Ethanol Ratio
on Ignition and Combustion of Ethanol-Gasoline Blend Spray in DISI Engine-Like
Condition. SAE Int. J. Fuels Lubr. 8(2):264-276, 2015, doi:10.4271/2015-01-0774.
DANIEL, R., TIAN, G., XU, H., WYSZYNSKI, M.L., WU, X. AND HUANG, Z. 2011.
Effect of spark timing and load on a DISI engine fuelled with 2,5-dimethylfuran. Fuel, 90,
449-458, doi:10.1016/j.fuel.2010.10.008.
242
DANIEL, R., XU, H., WANG, C., RICHARDSON, D. AND SHUAI, S. 2012a. Combustion
performance of 2,5-dimethylfuran blends using dual-injection compared to direct-injection in
a SI engine. Applied Energy, 98, 59-68, doi:http://dx.doi.org/10.1016/j.apenergy.2012.02.073
DANIEL, R., WANG, C., XU, H., TIAN, G. AND RICHARDSON, D. 2012b. Dual-
Injection as a Knock Mitigation Strategy Using Pure Ethanol and Methanol. SAE Int. J. Fuels
Lubr. 5(2):772-784, doi:10.4271/2012-01-1152.
DANIEL, R., WANG, C., XU, H. AND TIAN, G. 2012c. Effects of Combustion Phasing,
Injection Timing, Relative Air-Fuel Ratio and Variable Valve Timing on SI Engine
Performance and Emissions using 2,5-Dimethylfuran. SAE Int. J. Fuels Lubr. 5(2):855-866,
doi:10.4271/2012-01-1285.
DANIEL, R., TIAN, G., XU, H. AND SHUAI, S. 2012e. Ignition timing sensitivities of
oxygenated biofuels compared to gasoline in a direct-injection SI engine. Fuel. 99, 72-82,
doi:http://dx.doi.org/10.1016/j.fuel.2012.01.053.
DAVIS, G. AND BORGNAKKE, C. 1982. The Effect of In-Cylinder Flow Processes (Swirl,
Squish and Turbulence Intensity) on Engine Efficiency — Model Predictions. SAE Technical
Paper 820045, doi:10.4271/820045.
DELPHI 2012. Worldwide Emissions Standards - Passenger Cars and Light Duty Vehicles.
DENG, B., YANG, J. AND ZHANG, D. 2013a. The challenges and strategies of butanol
application in conventional engines: The sensitivity study of ignition and valve timing.
Applied Energy, 108, 248-260, doi:http://dx.doi.org/10.1016/j.apenergy.2013.03.018.
DENG, B., FU, J., ZHANG, D., YANG, J., FENG, R., LIU, J., LI, K. AND LIU, X. 2013b.
The heat release analysis of bio-butanol/gasoline blends on a high speed SI (spark ignition)
engine. Energy. 60, 230–241, doi:10.1016/j.energy.2013.07.055.
243
DIANA, S., GIGLIO, V., IORIO, B. AND POLICE, G. 1996. A Strategy to Improve the
Efficiency of Stoichiometric Spark Ignition Engines. SAE Technical Paper 961953,
doi:10.4271/961953.
DIRECTIVE 2009/28/EC. 2009. On the promotion of the use of energy from renewable
sources. Official Journal of the European Union; 2009.
EASTWOOD, P. 2008. Particulate Emissions from Vehicles, Chichester, John Wiley & Sons
Ltd.
ENDRES, H., NEUΒER, H. AND WURMS, R. 1992. Influence of Swirl and Tumble on
Economy and Emissions of Multi Valve SI Engines. SAE Technical Paper 920516,
doi:10.4271/920516.
EUROPEAN COUNCIL. 2016. Vehicle emissions in real driving conditions: Council gives
green light to second package. Available: http://www.consilium.europa.eu/en/press/press-
releases/2016/02/12-vehicle-emissions-in-real-driving-conditions-2nd-package/. Last
accessed 6th Apr 2016.
FENG, R., YANG, J., ZHANG, D., DENG, B., FU, J., LIU, J. AND LIU, X. 2013.
Experimental study on SI engine fuelled with butanol–gasoline blend and H2O addition.
Energy Conversion and Management. 74, 192–200, doi:10.1016/j.enconman.2013.05.021.
FLOCH, A., VAN FRANK, J. AND AHMED, A. 1995. Comparison of the Effects of Intake-
Generated Swirl and Tumble on Turbulence Characteristics in a 4-Valve Engine. SAE
Technical Paper 952457, doi:10.4271/952457.
DENG, B., FU, J., ZHANG, D., YANG, J., FENG, R., LIU, J., LI, K. AND LIU, X. 2013b.
The heat release analysis of bio-butanol/gasoline blends on a high speed SI (spark ignition)
engine. Energy. 60, 230–241, doi:10.1016/j.energy.2013.07.055.
244
FRENNET, A. AND BASTIN, J.-M. 1995. Catalysis and Automotive Pollution Control III.
Amsterdam: Elsevier.
FU, J., DENG, B., WANG, Y., YANG, J., ZHANG, D., XU, Z. AND LIU, J. 2014.
Numerical study and correlation development on laminar burning velocities of n-butanol, iso-
octane and their blends: Focusing on diluent and blend ratio effects. Fuel. 124, 102–112,
doi:10.1016/j.fuel.2014.01.092.
GILL, S., TURNER, D., TSOLAKIS, A. AND YORK, A. 2011. Understanding the Role of
Filtered EGR on PM Emissions. SAE Technical Paper 2011-01-2080, doi:10.4271/2011-01-
2080.
GALLONI, E., FONTANA, G. AND PALMACCIO, R. 2013. Effects of exhaust gas recycle
in a downsized gasoline engine. Applied Energy. 105, 99–107,
doi:10.1016/j.apenergy.2012.12.046.
GOV UK. 2011. Public attitudes towards climate change and the impact of transport: 2010.
Available:
https://www.gov.uk/government/uploads/system/uploads/attachment_data/file/8916/climate-
change-2010-report.pdf. Last accessed 27th February 2015.
GRAHAM, L. A., BELISLE, S. L. AND BAAS, C.-L. 2008. Emissions from light duty
gasoline vehicles operating on low blend ethanol gasoline and E85. Atmospheric
Environment. 42 (19), 4498–4516, doi:10.1016/j.atmosenv.2008.01.061.
GREEN CAR CONGRESS. 2006. Mercedes-Benz Premiers New Gasoline Direct Injection
System for More Power and Lower Fuel Consumption. Available:
http://www.greencarcongress.com/2006/02/mercedesbenz_pr.html. Last accessed 26th Feb
2015.
GREEN CAR CONGRESS. 2013. 2013 SAE International High Octane Fuels Symposium:
the potential for high octane fuels (Part 1). Available:
http://www.greencarcongress.com/2013/01/hofs-20130131.html. Last accessed 25th Feb
2015.
245
GU, X., HUANG, Z., CAI, J., GONG, J., WU, X. AND LEE, C.-F. 2012. Emission
characteristics of a spark-ignition engine fuelled with gasoline-n-butanol blends in
combination with EGR. Fuel, 93, 611-617, doi:10.1016/j.fuel.2011.11.040.
GUMBLETON, J., NIEPOTH, G. AND CURRIE, J. 1976. Effect of Energy and Emission
Constraints on Compression Ratio. SAE Technical Paper 760826, doi:10.4271/760826.
GUPTA, H. N. 2013. Fundamentals of Internal Combustion Engines. 2nd ed. Delhi, India:
Phi Learning.
HARADA, J., TOMITA, T., MIZUNO, H., MASHIKI, Z. AND ITO, Y. 1997. Development
of Direct Injection Gasoline Engine. SAE Technical Paper 970540, doi:10.4271/970540.
HE, B.-Q., WANG, J.-X., HAO, J.-M., YAN, X.-G. AND XIAO, J.-H. 2003. A study on
emission characteristics of an EFI engine with ethanol blended gasoline fuels. Atmospheric
Environment. 37 (7), 949–957, doi:10.1016/S1352-2310(02)00973-1.
HE, X., IRELAND, J., ZIGLER, B., RATCLIFF, M. A., KNOLL, K. E., ALLEMAN, T. L.
AND TESTER, J. T. 2010. The Impacts of Mid-level Biofuel Content in Gasoline on SIDI
Engine-out and Tailpipe Particulate Matter Emissions. SAE Technical Paper 2010-01-2125,
doi:10.4271/2010-01-2125.
HE, B.-Q., LIU, M.-B. AND ZHAO, H. 2015. Comparison of combustion characteristics of
n-butanol/ ethanol–gasoline blends in a HCCI engine. Energy Conversion and Management.
95, 101–109, doi:10.1016/j.enconman.2015.02.019.
HEDGE, M., WEBER, P., GINGRICH, J., ALGER, T. AND KHALEK, I. 2011. Effect of
EGR on Particle Emissions from a GDI Engine. SAE Int. J. Engines 4(1):650-666,
doi:10.4271/2011-01-0636.
HEFFEL, J. W. 2003. NOx emission and performance data for a hydrogen fueled internal
combustion engine at using exhaust gas recirculation. International Journal of Hydrogen
Energy. 28, 901–908, doi:10.1016/S0360-3199(02)00157-X.
HIROOKA, H., MORI, S. AND SHIMIZU, R. 2004. Effects of High Turbulence Flow on
Knock Characteristics. SAE Technical Paper 2004-01-0977, doi:10.4271/2004-01-0977.
246
HOFFMEYER, H., MONTEFRANCESCO, E., BECK, L., WILLAND, J., ZIEBART, F.
AND MAUSS, F. 2009. CARE – CAtalytic Reformated Exhaust Gases in Turbocharged
DISI-Engines, SAE Int. J. Fuels Lubr. 2(1):139-148, doi:10.4271/2009-01-0503.
HSIEH, W.-D., CHEN, R.-H., WU, T.-L. AND LIN, T.-H. 2002. Engine performance and
pollutant emission of an SI engine using ethanol–gasoline blended fuels. Atmospheric
Environment. 36 (3), 403–410, doi:10.1016/S1352-2310(01)00508-8.
HU, E., HU, X., WANG, X., XU, Y., DEARN, K. D. AND XU, H. 2012. On the
fundamental lubricity of 2,5-dimethylfuran as a synthetic engine fuel. Tribology
International, 55, 119–125, doi:http://dx.doi.org/10.1016/j.triboint.2012.06.005
HULL, A., GOLUBKOV, I., KRONBERG, B., MARANDZHEVA, T. AND VAN STAM, J.
2005. An alternative fuel for spark ignition engines. International Journal of Engine
Research, 7 , 203-214, doi:10.1243/14680874JER02504.
IKEYA, K., TAKAZAWA, M., YAMADA, T., PARK, S. AND TAGISHI, R. 2015. Thermal
Efficiency Enhancement of a Gasoline Engine. SAE Int. J. Engines 8(4), doi:10.4271/2015-
01-1263.
ISERMANN, R. 2014. Engine Modeling and Control- Modeling and Electronic Management
of Internal Combustion Engines. New York: Springer-Verlag.
JOHNSON, T. 2010. Review of CO2 Emissions and Technologies in the Road Transportation
Sector. SAE Int. J. Engines 3(1):1079-1098, doi:10.4271/2010-01-1276.
KALGHATGI, G. 2014. The outlook for fuels for internal combustion engines. International
Journal of Engine Research. Special Article, 1-17, doi: 10.1177/1468087414526189.
KARAVALAKIS, G., SHORT, D., HAJBABAEI, M., VU, D., VILLELA, M., RUSSELL,
R., DURBIN, T. AND ASA-AWUKU, A. 2013. Criteria Emissions, Particle Number
Emissions, Size Distributions, and Black Carbon Measurements from PFI Gasoline Vehicles
Fuelled with Different Ethanol and Butanol Blends. SAE Technical Paper 2013-01-1147,
doi:10.4271/2013-01-1147.
247
KARAVALAKIS, G., SHORT, D., VU, D., VILLELA, M., ASA-AWUKU, A. AND
DURBIN, T. D. 2014. Evaluating the regulated emissions, air toxics, ultrafine particles, and
black carbon from SI-PFI and SI-DI vehicles operating on different ethanol and iso-butanol
blends. Fuel. 128, 410–421, doi:10.1016/j.fuel.2014.03.016.
KAWABATA, Y., SAKONJI, T. AND AMANO, T. 1999. The Effect of NOx on Knock in
Spark-ignition Engines. SAE Technical Paper 1999-01-0572, doi:10.4271/1999-01-0572.
KESKIN, A. AND GÜRÜ, M. 2011. The Effects of Ethanol and Propanol Additions Into
Unleaded Gasoline on Exhaust and Noise Emissions of a Spark Ignition Engine. Energy
Sources, Part A: Recovery, Utilization, and Environmental Effects. 33 (23), 2194-2205, doi:
10.1080/15567030903530558.
KIM, S. AND KIM, S. 1995. Effects of Swirl and Spark Plug Shape on Combustion
Characteristic in a High Speed Single-Shot Visualized SI Engine. SAE Technical Paper
951003, doi:10.4271/951003.
KIM, M., LEE, S. AND KIM, W. 2006. Tumble Flow Measurements Using Three Different
Methods and its Effects on Fuel Economy and Emissions. SAE Technical Paper 2006-01-
3345, 2006, doi:10.4271/2006-01-3345.
KRAMER, F., SCHWARZ, C. AND WITT, A. 2000. Effect of Compression Ratio on the
Combustion of a Pressure Charged Gasoline Direct Injection Engine. SAE Technical Paper
2000-01-0250, doi:10.4271/2000-01-0250.
LATTIMORE, T., WANG, C., XU, H., WYSZYNSKI, M. L. AND SHUAI, S. 2016a.
Investigation of EGR Effect on Combustion and PM Emissions in a DISI Engine, Applied
Energy, 161, 2016, 256-267, doi:10.1016/j.apenergy.2015.09.080.
LEE, S., TONG, K., QUAY, B., ZELLO, J. AND SANTAVICCA, D. A. 2000. Effects of
Swirl and Tumble on Mixture Preparation During Cold Start of a Gasoline Direct-Injection
Engine. SAE Technical Paper 2000-01-1900, doi:10.4271/2000-01-1900.
248
LEE, K., BAE, C. AND KANG, K. 2007. The effects of tumble and swirl flows on flame
propagation in a four-valve S.I. engine. Applied Thermal Engineering. 27 (11-12), 2122–
2130, doi:10.1016/j.applthermaleng.2006.11.011.
LEWIS, A., AKEHURST, S., TURNER, J., PATEL, R. AND POPPLEWELL, A. 2014.
Observations on the Measurement and Performance Impact of Catalyzed vs. Non Catalyzed
EGR on a Heavily Downsized DISI Engine. SAE Int. J. Engines 7(1):458-467,
doi:10.4271/2014-01-1196.
LI, Y., ZHAO, H., PENG, Z., AND LADOMMATOS, N. 2001. Analysis of Tumble and
Swirl Motions in a Four-Valve SI Engine. SAE Technical Paper 2001-01-3555,
doi:10.4271/2001-01-3555.
LIIVA, P., VALENTINE, J., COBB, J., AND ACKER, W. 1992. Use of Multiple Pressure
Transducers to Find In-Cylinder Knock Location. SAE Technical Paper 922368,
doi:10.4271/922368.
LIU, H., LI, S., ZHENG, Z., XU, J. AND YAO, M. 2013. Effects of n -butanol, 2-butanol,
and methyl octynoate addition to diesel fuel on combustion and emissions over a wide range
of exhaust gas recirculation (EGR) rates. Applied Energy. 112, 246-256,
doi:10.1016/j.apenergy.2013.06.023.
MA, X., ZHANG, F., XU, H. AND SHUAI, S. 2014. Throttleless and EGR-controlled
stoichiometric combustion in a diesel–gasoline dual-fuel compression ignition engine, Fuel,
115, 765–777, doi:10.1016/j.fuel.2013.07.052.
MAJI, S., BABU, M. AND GUPTA, N. 2001. A Single Cylinder Engine Study of Power,
Fuel Consumption and Exhaust Emissions with Ethanol. SAE Technical Paper 2001-28-0029,
doi:10.4271/2001-28-0029.
249
MATSUSHITA, S., INOUE, T., NAKANISHI, K., OKUMURA, T. AND ISOGAI, K. 1985.
Effects of Helical Port with Swirl Control Valve on the Combustion and Performance of S. I.
Engine. SAE Technical Paper 850046, doi:10.4271/850046.
MEHTA, D., ALGER, T., HALL, M., MATTHEWS, R. D., AND NG, H. 2001. Particulate
Characterization of a DISI Research Engine using a Nephelometer and In-Cylinder
Visualization‖. SAE Technical Paper 2001-01-1976, doi:10.4271/2001-01-1976.
MITTAL, V., REVIER, B. AND HEYWOOD, J. 2007. Phenomena that Determine Knock
Onset in Spark-Ignition Engines. SAE Technical Paper 2007-01-0007, doi:10.4271/2007-01-
0007.
NAGAYAMA, I., ARAKI, Y. AND IIOKA, Y. 1977. Effects of Swirl and Squish on S.I.
Engine Combustion and Emission. SAE Technical Paper 770217, doi:10.4271/770217.
NIASS, T., AMER, A., XU, W., VOGEL, S., HORTMANN, K. K., ADOMEIT, P. AND
BRASSAT, A. 2012. Butanol Blending - a Promising Approach to Enhance the
Thermodynamic Potential of Gasoline - Part 1. SAE Int. J. Fuels Lubr. 5(1):265-273,
doi:10.4271/2011-01-1990.
NIEMAN, D., DEMPSEY, A., AND REITZ, R. 2012. Heavy-Duty RCCI Operation Using
Natural Gas and Diesel. SAE Int. J. Engines 5(2):270-285, doi:10.4271/2012-01-0379.
NISHIDA, M., INOUE, N., SUZUKI, H., AND KUMAGAI, S. 1988. Closed Loop Control
of the EGR Rate Using the Oxygen Sensor. SAE Technical Paper 880133,
doi:10.4271/880133.
OJAPAH, M., ZHAO, H. AND ZHANG, Y. 2014. Effects of Ethanol on Performance and
Exhaust Emissions from a DI Spark Ignition Engine with Throttled and Unthrottled
Operations. SAE Technical Paper 2014-01-1393, doi:10.4271/2014-01-1393.
250
OKAMOTO, K., ICHIKAWA, T., SAITOH, K., OYAMA, K., HIRAYA, K. AND
URUSHIHARA, T. 2003. Study of Antiknock Performance Under Various Octane Numbers
and Compression Ratios in a DISI Engine. SAE Technical Paper 2003-01-1804,
doi:10.4271/2003-01-1804.
OWEN, K. AND COLEY, T. 1995. Automotive Fuels Reference Book, Warrendale, Society
of Automotive Engineers.
PAN, M., SHU, G., WEI, H., ZHU, T., LIANG, Y. AND LIU, C. 2014. Effects of EGR,
compression ratio and boost pressure on cyclic variation of PFI gasoline engine at WOT
operation. Applied Thermal Engineering, 64, 491-498,
doi:10.1016/j.applthermaleng.2013.11.013.
PARSONS, D., AKEHURST, S. AND BRACE, C. 2015. The potential of catalysed exhaust
gas recirculation to improve high-load operation in spark ignition engines. International
Journal of Engine Research. 16 (4), 592-605, doi:10.1177/1468087414554628.
RAMADHAS, A. S. 2011. Alternative Fuels for Transportation. Boca Raton, FL: CRC Press.
RATNAK, S., KATORI, K., KUSAKA, J., DAISHO, Y., YOSHIMURA, K. AND
KENJIRO, N. 2015. Computational Study to Improve Thermal Efficiency of Spark Ignition
Engine. SAE Technical Paper 2015-01-0011, doi:10.4271/2015-01-0011.
REITZE JR., A. W. 2001. Air Pollution Control Law: Compliance & Enforcement.
Washington D. C.: Environmental Law Institute.
251
RHODES, D. AND KECK, J. 1985. Laminar Burning Speed Measurements of Indolene-Air-
Diluent Mixtures at High Pressures and Temperatures. SAE Technical Paper 850047,
doi:10.4271/850047.
ROTH, D., KELLER, P. AND BECKER, M. 2010. Requirements of External EGR Systems
for Dual Cam Phaser Turbo GDI Engines. SAE Technical Paper 2010-01-0588,
doi:10.4271/2010-01-0588.
SIOKOS, K., KOLI, R., PRUCKA, R., SCHWANKE, J. AND MIERSCH, J. 2015.
Assessment of Cooled Low Pressure EGR in a Turbocharged Direct Injection Gasoline
Engine. SAE Int. J. Engines 8(4), doi:10.4271/2015-01-1253.
SONG, D., JIA, N., GUO, X., MA, X., MA, Z., GAO, D., LI, K., LAI, H. AND ZHANG, C.
2014. Low Pressure Cooled EGR for Improved Fuel Economy on a Turbocharged PFI
Gasoline Engine. SAE Technical Paper 2014-01-1240, doi:10.4271/2014-01-1240.
252
STONE, R., CHEN, L., HINTON, N. LEACH, F. AND XU, F. 2012. GDI Engine Operation
with Ethanol/Gasoline Blends and Aqueous Ethanol. Journal of Automotive Safety and
Energy. 3 (3), 257-264, doi:10.3969/j.issn.1674-8484.2012.03.009.
STOREY, J., BARONE, T., NORMAN, K. AND LEWIS, S. 2010. Ethanol Blend Effects On
Direct Injection Spark-Ignition Gasoline Vehicle Particulate Matter Emissions. SAE Int. J.
Fuels Lubr. 3(2):650-659, doi:10.4271/2010-01-2129.
SU, J., XU, M., LI, T., GAO, Y. AND WANG, J. 2014. Combined effects of cooled EGR
and a higher geometric compression ratio on thermal efficiency improvement of a downsized
boosted spark-ignition direct-injection engine. Energy Conversion and Management. 78, 65–
73, doi:10.1016/j.enconman.2013.10.041.
TAKAHASHI, D., NAKATA, K., YOSHIHARA, Y., OHTA, Y. AND NISHIURA, H. 2015.
Combustion Development to Achieve Engine Thermal Efficiency of 40% for Hybrid
Vehicles. SAE Technical Paper 2015-01-1254, doi:10.4271/2015-01-1254.
TAKAKI, D., TSUCHIDA, H., KOBARA, T., AKAGI, M., TSUYUKI, T. AND
NAGAMINE, M. 2014. Study of an EGR System for Downsizing Turbocharged Gasoline
Engine to Improve Fuel Economy. SAE Technical Paper 2014-01-1199, doi:10.4271/2014-
01-1199.
TERAJI, A., KAKUHO, A., TSUDA, T. AND HASHIZUME, Y. 2009. A Study of the
Knocking Mechanism in Terms of Flame Propagation Behavior Based on 3D Numerical
Simulations. SAE Int. J. Engines 2(1):666-673, doi:10.4271/2009-01-0699.
TIAN, G., LI, H, XU, H., LI, Y, AND SATISH M.R. 2010. Spray Characteristics Study of
DMF Using Phase Doppler Particle Analyzer. SAE 2010-01-1505, doi:10.4271/2010-01-
1505.
TOMITA, M., IWAKIRI, Y., SAKAI, E., URUSHIHARA, T., INOUE, R. AND KOJIMA,
K. 1996. Effects of Gas Flow and Mixture Properties on Engine-Out HC Emissions. SAE
Technical Paper 961952, doi:10.4271/961952.
253
TURNER, D. 2010. The combustion and emissions performance of fuel blends in modern
combustion systems. PhD Thesis, University of Birmingham.
TURNER, D., XU, H., CRACKNELL, R. F., NATARAJAN, V. AND CHEN, X. 2011.
Combustion performance of bio-ethanol at various blend ratios in a gasoline direct injection
engine. Fuel. 90, 1999-2006, doi:10.1016/j.fuel.2010.12.025.
URUSHIHARA, T., MURAYAMA, T., TAKAGI, Y. AND LEE, K. 1995. Turbulence and
Cycle-by-Cycle Variation of Mean Velocity Generated by Swirl and Tumble Flow and Their
Effects on Combustion. SAE Technical Paper 950813, doi:10.4271/950813.
VAROL, Y., ÖNER, C., ÖZTOP, H. F. AND ALTUN, Ş. 2014. Comparison of Methanol,
Ethanol, or n-Butanol Blending with Unleaded Gasoline on Exhaust Emissions of an SI
Engine. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects. 36 (9),
938-948, doi:10.1080/15567036.2011.572141.
VUK, C. AND VANDER GRIEND, S. 2013. Fuel Property Effects on Particulates In Spark
Ignition Engines. SAE Technical Paper 2013-01-1124, doi:10.4271/2013-01-1124.
WALLNER, T., ICKES, A., LAWYER, K., MIERS, S., NABER, J., ERTL, D.,
WILLIAMSON, R., MIERS, S. A. AND NABER, J. D. 2013. Blend Ratio Optimization of
Fuels Containing Gasoline Blendstock, Ethanol, and Higher Alcohols (C3-C6): Part II -
Blend Properties and Target Value Sensitivity. SAE International 2013-01-1126,
doi:10.4271/2013-01-1126.
WANG, C., XU, H., DANIEL, R., GHAFOURIAN, A., HERREROS, J. M., SHUAI, S.
AND MA, X. 2013. Combustion characteristics and emissions of 2-methylfuran compared to
2,5-dimethylfuran, gasoline and ethanol in a DISI engine. Fuel, 103, 200-211,
doi:http://dx.doi.org/10.1016/j.fuel.2012.05.043.
WANG, C., XU, H., HERREROS, J. M., WANG, J. AND CRACKNELL, R. 2014a. Impact
of fuel and injection system on particle emissions from a GDI engine. Applied Energy, 132,
178–191, doi:http://dx.doi.org/10.1016/j.apenergy.2014.06.012.
254
WANG, C. 2014b. Combustion and emissions of a direct injection gasoline engine using
biofuels. PhD Thesis, University of Birmingham.
WANG, C., XU, H., HERREROS, J. M., LATTIMORE, T. AND SHUAI, S. 2014c. Fuel
Effect on Particulate Matter Composition and Soot Oxidation in a Direct-Injection Spark
Ignition (DISI) Engine. Energy & Fuels, 28, 2003-2012, doi:10.1021/ef402234z.
WEI, H., ZHU, T., SHU, G., TAN, L. AND WANG, Y. 2012. Gasoline engine exhaust gas
recirculation – A review. Applied Energy. 99, 534–544, doi:10.1016/j.apenergy.2012.05.011.
YANG, J. AND ANDERSON, R.W. 1998. Fuel Injection Strategies to Increase Full-Load
Torque Output of a Direct-Injection SI Engine. SAE Technical Paper 980495,
doi:10.4271/980495.
YEJIAN, Q., CHENGJI, Z., HONGMING, X., JIAN, T. AND QIN, T. 2007. Experimental
Investigation of Using Ethanol-gasoline in Electronic Control Gasoline-injection Engine.
HKIE Transactions. 14 (2), 26-30, doi:10.1080/1023697X.2007.10668074.
YUAN, H., FOONG, T., CHEN, Z., YANG, Y., BREAR, M., LEONE, T. AND
ANDERSON, J. E. 2015. Modeling of Trace Knock in a Modern SI Engine Fuelled by
Ethanol/Gasoline Blends. SAE Technical Paper 2015-01-1242, doi:10.4271/2015-01-1242.
ZACCARDI, J., DUVAL, L. AND PAGOT, A. 2009. Development of Specific Tools for
Analysis and Quantification of Pre-ignition in a Boosted SI Engine. SAE Int. J. Engines
2(1):1587-1600, doi:10.4271/2009-01-1795.
ZHANG, Z., WANG, T., JIA, M., WEI, Q., MENG, X. AND SHU, G. 2014a. Combustion
and particle number emissions of a direct injection spark ignition engine operating on
ethanol/gasoline and n-butanol/gasoline blends with exhaust gas recirculation. Fuel, 130,
177-188, doi:http://dx.doi.org/10.1016/j.fuel.2014.04.052.
ZHANG, Z., ZHANG, H., WANG, T. AND JIA, M. 2014b. Effects of tumble combined with
EGR (exhaust gas recirculation) on the combustion and emissions in a spark ignition engine
at part loads. Energy. 65, 18–24, doi:10.1016/j.energy.2013.11.062.
ZHANG, J., NITHYANANDAN, K., LI, Y., LEE, C.-F. AND HUANG, Z. 2015.
Comparative Study of High-Alcohol-Content Gasoline Blends in an SI Engine. SAE
Technical Paper 2015-01-0891, doi:10.4271/2015-01-0891.
255
ZHONG, L., MUSIAL, M., REESE, R. AND BLACK, G. 2013. EGR Systems Evaluation in
Turbocharged Engines. SAE Technical Paper 2013-01-0936, doi:10.4271/2013-01-0936.
256