Sei sulla pagina 1di 322

A Study of the Knock Limits of

Liquefied Petroleum Gas (LPG) in


Spark-Ignition Engines
Kai J. Morganti

Submitted in total fulfilment of the requirements of the degree of

Doctor of Philosophy

Department of Mechanical Engineering


T HE U NIVERSITY OF M ELBOURNE

December 2013

Produced on archival quality paper.


Copyright
c 2013 Kai J. Morganti

All rights reserved. No part of the publication may be reproduced in any form by print,
photoprint, microfilm or any other means without written permission from the author.
Abstract

A substantial increase in the use of alternative transport fuels is considered by some to be


an important part of our response to climate change. Liquefied Petroleum Gas (LPG) is
one such fuel that shows significant potential to improve the efficiency and greenhouse gas
emissions of conventional spark-ignition engines. However, there is still some uncertainty
as to the best use of LPG in spark-ignition engines. This uncertainty can largely be
attributed to variations in the composition of LPG, along with the susceptibility of these
different mixtures to so-called autoignition. Often, this undesirable form of combustion
leads to potentially damaging rates of in-cylinder pressure rise, and therefore should be
avoided. This requirement places an upper limit on the efficiency of the spark-ignition
engine.
This thesis therefore aims to develop a fundamental understanding of the mecha-
nisms responsible for the autoignition of LPG mixtures under conditions relevant to
spark-ignition engines. A comprehensive experimental study of the susceptibility of
different LPG mixtures to autoignition is first presented. This utilises the standard ASTM
Research and Motor methods for liquid fuels, which are adapted to enable the autoignition
propensity of different LPGs to be quantified in terms of the so-called Research and Motor
octane numbers (RON and MON respectively). The engine experiments are then examined
numerically using a multi-zone combustion model that incorporates detailed chemical
kinetics. This model is calibrated and validated using empirical data, and is used to
investigate the key kinetic pathways leading to the autoignition of different LPG mixtures.
The work presented in this thesis first demonstrates that the linear blending of octane
numbers appears to be reasonable for most LPGs. It is also shown that almost all LPGs
have higher RONs than both standard and premium gasolines. Importantly, imposition
of the existing MON and vapour pressure requirements specified in the Australian LPG
fuel standard further increases this advantage over all gasolines. This suggests that the
existing fuel standards unnecessarily restrict the composition of commercial LPGs, given
the widespread use of LPG in retro-fitted gasoline vehicles at present.
The modelling results indicate that an accurate model of LPG autoignition in a spark-
ignition engine needs to include several key phenomena. In particular, careful modelling of
the in-cylinder heat transfer and residual nitric oxide concentration should be performed,
in addition to the flame propagation modelling and autoignition chemistry. When these
key phenomena are included in the model, the autoignition of most LPG mixtures could
be modelled to within 1.0 crank angle degree (0.28 ms) of experiment.

iii
Declaration
This is to certify that

1. the thesis comprises only my original work towards the degree of Doctor of Philoso-
phy except where indicated in the Preface,

2. due acknowledgement has been made in the text to all other material used,

3. the thesis is fewer than 100,000 words in length, exclusive of tables, maps, bibliogra-
phies and appendices.

Kai J. Morganti, December 2013

v
Acknowledgements

I’d like to express my gratitude to several people that have provided me with immense
support throughout this project. To my supervisors, Michael Brear and Gabriel da Silva,
thank you for your wealth of knowledge, expertise and personal interest in the undertak-
ings of this project. Thanks also to my honorary supervisors, Harry Watson, Yi Yang and
Fred Dryer. Thank you for your support, friendship and the skills that each of you have
assisted me to develop.

I’d also like to thank our administration staff within the department, particularly Jan
May, Carol Barrie, Monica Pater and Eileen Doufas-Shea. Without your valuable work,
projects like these simply would not happen. Similarly, I’d like to thank Al Knox and Don
Halpin for their valuable support with the technical aspects of the project, along with their
friendship over the last three years.

To my friends within the Thermodynamics Group, particularly Ashley Wiese, Mohsen


Talei, Denis Andianov, Payman Abbasi Atibeh, Peter Dennis, Peter Hield, Matthew Blom,
Bishoy Alfons, Tim Broomhead, Matthew Jeppesen, Josh Lacey, Farzad Poursadegh, Ali
Haghiri and Rob Dingli- thank you for making the last three years such an enjoyable
experience.

Foremost, I wish to extend thanks to my parents, Adrian and Anna, for their continued
support throughout this long and often arduous process. I could never have reached this
point without the encouragement and opportunity you have provided me with over the
last 27 years.

vii
Peer Reviewed Publications
K.J. Morganti, T.M Foong, M.J. Brear, G. da Silva, Y. Yang and F.L. Dryer, “The Autoignition
of Liquefied Petroleum Gas (LPG) in Spark-Ignition Engines,” submitted to International
Symposium on Combustion, San Francisco, USA, 2014.

K.J. Morganti, T.M Foong, M.J. Brear, G. da Silva, Y. Yang and F.L. Dryer, “Design and
Analysis of a Modified CFR Engine for the Octane Rating of Liquefied Petroleum Gases
(LPG),” submitted to SAE World Congress, Detroit, USA, 2014.

K.J. Morganti, T.M Foong, M.J. Brear, G. da Silva, Y. Yang and F.L. Dryer, “The Autoigni-
tion of Propane and n-Butane During Octane Rating: Engine Experiments and Detailed
Chemical Kinetic Modelling,” Proc. Aust. Combust. Sym., Nov. 6 - 8, 2013.

K.J. Morganti, T.M Foong, M.J. Brear, G. da Silva, Y. Yang and F.L. Dryer, “The Research
and Motor Octane Numbers of Liquefied Petroleum Gas (LPG),” Fuel 108, pp.797-811, 2013.

K.J. Morganti, T.M Foong, M.J. Brear, G. da Silva and F.L. Dryer, “A Comparison of
Reference Engine Autoignition with Detailed Kinetic Modelling for a range of Primary
Reference Fuels,” Proc. Aust. Combust. Sym., Nov. 29 - Dec 1, 2011.

ix
There’s a time to be Daniel Boone, and a time to be a plumber.
This one called for the latter...

MacGyver

xi
Contents

1 Introduction 1
1.1 The transport sector and climate change . . . . . . . . . . . . . . . . . . . . 1
1.2 Alternative transport fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Fuel suitability for spark-ignition engines . . . . . . . . . . . . . . . . . . . 3
1.4 Thesis layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2 Literature Review 9
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 What is Liquefied Petroleum Gas? . . . . . . . . . . . . . . . . . . . . . . . 9
2.2.1 The history of LPG . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.2 The use of LPG as a transport fuel . . . . . . . . . . . . . . . . . . . 11
2.2.3 LPG composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.4 LPG fuel quality standards . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Knock in spark-ignition engines . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.1 Autoignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3.2 Pre-flame kinetic activity . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.3 Effect of residual combustion products on autoignition . . . . . . . 20
2.4 Test methods for the knock characteristics of spark-ignition engine fuels . 25
2.4.1 The Research and Motor methods . . . . . . . . . . . . . . . . . . . 25
2.4.2 The Motor (LP) method . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4.3 The Methane number . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4.4 Octane number blending in multi-component fuels . . . . . . . . . 31
2.4.5 The relevance of octane number to modern production engines . . 33
2.5 The knock characteristics of Liquefied Petroleum Gas . . . . . . . . . . . . 34
2.5.1 Research and Motor octane number data . . . . . . . . . . . . . . . 34
2.5.2 Motor octane number by gas chromatography . . . . . . . . . . . . 35
2.6 Modelling autoignition in combustion engines . . . . . . . . . . . . . . . . 38

xiii
2.6.1 Perfectly stirred reactor models with chemical kinetics . . . . . . . 39
2.6.2 Multi-zone engine models with chemical kinetics . . . . . . . . . . 40
2.6.3 Use of multi-zone models to study autoignition . . . . . . . . . . . 43
2.6.4 Detailed chemical kinetic mechanisms relevant to LPG . . . . . . . 44
2.7 Summary and research questions . . . . . . . . . . . . . . . . . . . . . . . . 46

3 Experimental Methods 49
3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.2 CFR Engine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.3 Dynamometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.4 Gaseous fuel preparation and delivery system . . . . . . . . . . . . . . . . 55
3.4.1 Storage cylinder heating . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.4.2 Leak detection system . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.4.3 Fuel flow measurement calibration verification . . . . . . . . . . . . 60
3.5 Standard test methods for octane number . . . . . . . . . . . . . . . . . . . 61
3.5.1 Engine Fit-for-Use procedure . . . . . . . . . . . . . . . . . . . . . . 63
3.5.2 Liquid reference and standardisation fuel preparation . . . . . . . . 63
3.5.3 Modified test method for gaseous fuels . . . . . . . . . . . . . . . . 64
3.6 In-cylinder pressure measurement . . . . . . . . . . . . . . . . . . . . . . . 66
3.6.1 Acquiring in-cylinder pressure data with autoignition suppressed 67
3.7 Measurement of the engine-out exhaust emissions . . . . . . . . . . . . . . 69
3.8 Measurement of the relative air-fuel ratio . . . . . . . . . . . . . . . . . . . 71

4 The Research and Motor Octane Numbers of Liquefied Petroleum Gas 73


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.2 Pure constituents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.3 Binary mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.4 An octane number model for LPG . . . . . . . . . . . . . . . . . . . . . . . 75
4.4.1 Experimental design and statistical analysis . . . . . . . . . . . . . 77
4.4.2 Regression analysis and response surface fitting . . . . . . . . . . . 78
4.4.3 Model validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.5 RONs and MONs for up to ternary mixtures . . . . . . . . . . . . . . . . . 87
4.5.1 Fuel sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.6 Validity of the EN 589 and ASTM D2598 MON calculation methods . . . . 92

xiv
4.6.1 Limitations of the MON calculation methods . . . . . . . . . . . . . 94
4.7 RONs of LPG satisfying existing fuel standards . . . . . . . . . . . . . . . . 95
4.7.1 Australian Fuel Determination (Autogas) . . . . . . . . . . . . . . . 95
4.7.2 European EN 589 standard . . . . . . . . . . . . . . . . . . . . . . . 95
4.7.3 United States Propane HD-5 . . . . . . . . . . . . . . . . . . . . . . 99
4.8 Implications for LPG fuel quality standards . . . . . . . . . . . . . . . . . . 100
4.9 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

5 Combustion Modelling and Analysis 103


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.2 In-cylinder pressure data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.2.1 Raw pressure traces with autoignition . . . . . . . . . . . . . . . . . 104
5.2.2 Criterion for selection of representative autoigniting pressure traces 106
5.2.3 Raw pressure traces without autoignition . . . . . . . . . . . . . . . 109
5.3 Engine-out emissions with autoignition suppressed . . . . . . . . . . . . . 112
5.4 Two-zone engine model without autoignition . . . . . . . . . . . . . . . . . 117
5.4.1 Model formulation and calibration . . . . . . . . . . . . . . . . . . . 119
5.4.2 Model validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
5.5 Analysis of modelling results . . . . . . . . . . . . . . . . . . . . . . . . . . 137
5.5.1 Mass fraction burned histories during octane rating . . . . . . . . . 137
5.5.2 Mass fraction burned at autoignition . . . . . . . . . . . . . . . . . . 137
5.5.3 Unburned zone temperature histories . . . . . . . . . . . . . . . . . 139
5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

6 Modelling Autoignition during Octane Rating using Chemical Kinetics 145


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
6.2 Two-zone engine model with autoignition . . . . . . . . . . . . . . . . . . . 146
6.2.1 Model formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.2.2 Chemical kinetic mechanisms . . . . . . . . . . . . . . . . . . . . . . 151
6.3 Effect of heat losses on autoignition . . . . . . . . . . . . . . . . . . . . . . . 154
6.4 Effect of humidity on autoignition . . . . . . . . . . . . . . . . . . . . . . . 154
6.5 Effect of carbon monoxide (CO) on autoignition . . . . . . . . . . . . . . . 156
6.6 Effect of unreacted and partially reacted hydrocarbons on autoignition . . 159
6.7 Effect of nitrogen oxides (NOx ) on autoignition . . . . . . . . . . . . . . . . 162

xv
6.7.1 Constant-volume ignition delay . . . . . . . . . . . . . . . . . . . . 162
6.7.2 Effect of nitric oxide (NO) on autoignition . . . . . . . . . . . . . . 164
6.7.3 Fuel oxidation in the presence of nitric oxide . . . . . . . . . . . . . 169
6.7.4 Effect of nitrogen dioxide (NO2 ) on autoignition . . . . . . . . . . . 174
6.8 Pre-flame kinetic activity within the unburned zone . . . . . . . . . . . . . 176
6.8.1 Fuel consumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
6.8.2 Unburned zone temperature increase prior to autoignition . . . . . 176
6.9 Further analysis of propylene autoignition . . . . . . . . . . . . . . . . . . 178
6.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

7 Conclusions and recommendations for future research 187


7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
7.2 Recommendations for future research . . . . . . . . . . . . . . . . . . . . . 191

A Research and Motor octane number data 227

B Empirical data used in the GT-Power model 229


B.1 Intake and exhaust valve lift profiles . . . . . . . . . . . . . . . . . . . . . . 229
B.2 Intake and exhaust valve discharge coefficient data . . . . . . . . . . . . . 230

C Calculation of the ideal products of combustion 233


C.1 Stoichiometric or fuel-lean combustion (φ ≤ 1) . . . . . . . . . . . . . . . . 233
C.2 Fuel-rich combustion (φ > 1) . . . . . . . . . . . . . . . . . . . . . . . . . . 235

D Speciated exhaust hydrocarbon measurements for propane 239

E Chemical kinetic mechanism 243

xvi
Nomenclature

A area [m2 ]
ALPGA Australian Liquefied Petroleum Gas Association
ANOVA Analysis of Variance
ASTM American Society for Testing and Materials
ABDC after bottom dead centre [degrees]
B cylinder bore [m]
BBDC before bottom dead centre [degrees]
BOP basic operator panel
ATDC after top dead centre [degrees]
(A/F) air-fuel ratio
(A/F)stoich stoichiometric air-fuel ratio
BDC bottom dead centre
BTDC before top dead centre [degrees]
CAD crank angle degrees
CARB California Air Resources Board
CARS Coherent anti-Stokes Raman spectroscopy
CFR Cooperative Fuel Research
CR compression ratio
DoE Design of Experiment
ei error term for observation i [octane number]
EGR exhaust gas recirculation
EOC end of combustion [degrees]
EVO exhaust valve opening [degrees]
F F-statistic
FCC Fluid Catalytic Cracker
GHG Greenhouse gas
GPA Gas Processors Association
hc convective heat transfer coefficient [W/m2 K]
HCCI Homogeneous Charge Compression Ignition
IMEP Indicated mean effective pressure [bar]
IP Institute of Petroleum
IPCC Intergovernmental Panel on Climate Change
ISO International Organisation for Standardisation
IVC intake valve closure [degrees]
KLCR knock limited compression ratio
KLSA knock limited spark advance [degrees]
KI knock Intensity [divisions]

xvii
Kp equilibrium constant
LFL lower flammability limit [%]
LPG Liquefied Petroleum Gas
LP Liquefied Petroleum
(mass/mass) mass fraction
m mass [kg]
ṁ mass flow rate [kg/s]
Mi Motor octane number blend factor for species i [octane number]
MFB Mass fraction burned [%]
MFC Mass Flow Controller
(mol/mol) mole fraction
MON Motor octane number
MN Methane Number
MSDS Material Safety Data Sheet
MSE mean square error
N total number of observations
n polytropic index
NDIR Nondispersive Infrared Detector
NGPA Natural Gasoline Processor Association
OI octane index
ON octane number
p pressure [bar or kPa]
p statistical significance threshold
PLM Professional Lambda Meter
ppm parts per million
PLM Professional Lambda Meter
PRF Primary Reference Fuel
PSR perfectly stirred reactor
q number of constituents in mixture
Q HV, i lower (net) heating value of species i [MJ/kg]
R2adj adjusted coefficient of determination
RON Research octane number
RSS Residual sum of squares
rpm engine speed [revolutions per minute]
S Fuel sensitivity [octane number]
SE standard error [octane number or crank angle degrees]
SOC Start of combustion [degrees]
S̄ p mean piston speed [m/s]
SI spark-ignition
SKI Standard Knock Intensity
SSE sum of squared errors
sub-CCR sub-critical compression ratio
TDC top dead centre
TEL tetraethyllead
TML tetramethyllead

xviii
TRF Toluene Reference Fuel
TSF Toluene Standardisation Fuel
T temperature [K]
Tb burned gas temperature [K]
Tu unburned gas temperature [K]
Tgas bulk gas temperature [K]
Twall cylinder wall temperature [K]
U internal energy [J]
Ub burned zone internal energy [J]
Uu unburned zone internal energy [J]
UEGO universal exhaust gas oxygen
UHCs Unburned hydrocarbons
US United States
V volume [m3 ]
Vb burned gas volume [m3 ]
Vu unburned gas volume [m3 ]
Vs swept volume [m3 ]
Vcyl, IVC cylinder volume at intake valve closure [m3 ]
Vr f , IVC residual fraction volume at intake valve closure [m3 ]
(vol/vol) volume fraction
WOT wide open throttle
w mean gas velocity [m/s]
Xj mole fraction of species j
yi measured octane number for observation i [octane number]
ŷi predicted octane number for observation i [octane number]
λ relative air-fuel ratio
ηc combustion efficiency [%]
ηv volumetric efficiency [%]
χi concentration of species i [ppm]
χi, IVC concentration of species i at intake valve closure [ppm]
χi, products concentration of species i within the combustion products [ppm]
β partial regression coefficient
φ equivalence ratio
ζ Woschni convective heat transfer coefficient multiplier

xix
List of Figures

1.1 Research and Motor octane numbers for various pure hydrocarbons, to-
gether with commercial gasolines meeting the requirements of the Aus-
tralian fuel standards. Data has been taken from several sources (Federal
Register of Legislative Instruments, 2001; Kubesh et al., 1992; Boldt, 1967;
American Petroleum Institute, 1958). . . . . . . . . . . . . . . . . . . . . . . 5

2.1 Structural formulas for propane, propylene, n-butane and iso-butane. . . . 15


2.2 High-speed schlieren cinematography showing the sequence of events
leading to autoignition in an optical access spark-ignition engine (Spicher
et al., 1991). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3 Cylinder pressure histories for (a) normal combustion, (b) medium or trace
knock and (c) heavy knock. Data was acquired in a single-cylinder engine
operated at 4000 rpm and wide open throttle (Douaud and Eyzat, 1977). . 19
2.4 Schematic diagram of the post-combustion processes by which hydrocar-
bons desorb and exit the cylinder during (a) the exhaust blowdown process;
(b) the exhaust stroke; (c) the end of the exhaust stroke (Lavoie et al., 1978). 22
2.5 Non-dimensional hydrocarbon concentration versus residence time for
methane (CH4 ), ethane (C2 H6 ), ethylene (C2 H4 ), propane (C3 H8 ) and propy-
lene (C3 H6 ) in the presence of 20 ppm of nitric oxide. Data was acquired in
a plug flow reactor at a temperature of 800 K (Hori et al., 1998). . . . . . . 24
2.6 The Cooperative Fuel Research octane rating unit (Balis, 2013). . . . . . . . 26
2.7 Schematic diagram of the gaseous fuel carburettor system used in the Motor
(LP) test method (Balis, 2012). . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.8 Linear and non-linear octane number blending between two hydrocarbons,
X1 and X2 . Segment (a) indicates synergistic blending, segment (b) indicates
linear blending, and segment (c) indicates antagonistic blending (Ghosh et al.,
2006). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.9 Estimated variation in the average value of K for spark-ignition engines
since 1932 (Mittal and Heywood, 2009). . . . . . . . . . . . . . . . . . . . . 34
2.10 Comparison between the MONs computed using the EN 589 and ASTM
D2598 standards, and the experimental octane data reported by Boldt (Boldt,
1967). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

xxi
2.11 The standard two-zone combustion model framework. . . . . . . . . . . . 41
2.12 Simulated autoignition onset for PRF40, PRF60 and PRF80 using a two-zone
model with chemical kinetics (Hajireza et al., 1998). . . . . . . . . . . . . . 44

3.1 CFR engine test facility layout. . . . . . . . . . . . . . . . . . . . . . . . . . 50


3.2 CFR F1/F2 octane rating engine without the gaseous fuel supply system
fitted. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.3 Gaseous fuel preparation and delivery system used in the experimental work. 56
3.4 Modifications performed to the standard intake system to enable operation
on gaseous fuels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.5 Piping and instrumentation diagram for the gaseous fuel preparation and
delivery system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.6 The Coriolis flow meter used to verify the calibration of the MFCs at the
commencement and completion of the experimental work. . . . . . . . . . 60
3.7 Measured fuel flow rates using the MFCs versus the Coriolis flow meter for
the four pure constituents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.8 Kistler 6125C piezoelectric pressure transducer used to acquire in-cylinder
pressure data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.9 Hardware used to add dilute TEL to the gaseous fuels. . . . . . . . . . . . 68
3.10 Measured in-cylinder pressure data for propane and n-butane at the Re-
search method SKI test condition, with and without TEL. Data is presented
for 300 consecutive cycles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

4.1 Experimentally determined RONs and MONs for binary mixtures of propane,
propylene, n-butane and iso-butane together with data reported by Boldt
(Boldt, 1967). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.2 The augmented {3,2} simplex lattice design used for the mixtures in this
study. Mixtures are presented on a molar basis. . . . . . . . . . . . . . . . . 78
4.3 Measured octane number versus predicted octane number for the design
point data, based on the regression models in Equations 4.6 and 4.7. . . . . 82
4.4 Design and check point locations in ternary factor space. Mixtures are
presented on a molar basis. . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.5 Measured octane number versus predicted octane number for the check
point data, based on the regression models in Equations 4.6 and 4.7. . . . . 83
4.6 Measured octane number versus predicted octane number for quaternary
mixtures, based on the regression models in Equations 4.6 and 4.7. . . . . 84

xxii
4.7 Predicted octane number versus error terms for all experimental data ac-
quired in this study, based on the regression models in Equations 4.6 and
4.7. The reproducibility limits for the standard test methods are provided
where available (ASTM International, 2012a,b). . . . . . . . . . . . . . . . . 86
4.8 Contour diagrams of RON for ternary mixtures of propane, propylene,
n-butane and iso-butane generated using the regression model in Equation
4.6. Data is presented on a molar basis. . . . . . . . . . . . . . . . . . . . . . 88
4.9 Contour diagrams of MON for ternary mixtures of propane, propylene,
n-butane and iso-butane generated using the regression model in Equation
4.7. Data is presented on a molar basis. . . . . . . . . . . . . . . . . . . . . . 89
4.10 Contour diagrams of fuel sensitivity for ternary mixtures of propane, propy-
lene, n-butane and iso-butane generated using the regression model in
Equation 4.11. Data is presented on a molar basis. . . . . . . . . . . . . . . 91
4.11 Measured MON versus predicted MON obtained using the EN 589 (Eu-
ropean Committee for Standardisation, 2008) and ASTM D2598 (ASTM
International, 2007a) calculation methods. . . . . . . . . . . . . . . . . . . . 93
4.12 Contour diagrams of RON for ternary mixtures of propane, propylene,
n-butane and iso-butane that comply with the MON and vapour pressure
requirements of the Australian Fuel Determination (Autogas) 2003 (Federal
Register of Legislative Instruments, 2003). Data is presented on a molar basis. 96
4.13 Contour diagrams of RON for ternary mixtures of propane, propylene,
n-butane and iso-butane that comply with the MON and vapour pressure
requirements of the European EN 589 standard (Grade D) (European Com-
mittee for Standardisation, 2008). Data is presented on a molar basis. . . . 98

5.1 Raw measured in-cylinder pressure traces and logarithmic pressure-volume


diagrams for propane, propylene, n-butane and iso-butane. Autoignition
occurs for all fuels, and is denoted for propane. Data is presented for the
Research method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . 105
5.2 Measured variation in the autoignition onset location for propane over 900
consecutive cycles. Data is presented for the Research method SKI test
condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.3 Measured raw in-cylinder pressure traces for propane, propylene, n-butane
and iso-butane. Data is presented for the cycles corresponding to the least
and most advanced combustion development, together with the represen-
tative cycle for each operating condition. All data was acquired at the
Research method SKI test condition, from a batch of 900 consecutive cycles. 108
5.4 Representative autoigniting and non-autoigniting pressure traces for propane,
propylene, n-butane and iso-butane. Data is presented for the Research
method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

xxiii
5.5 Representative autoigniting and non-autoigniting logarithmic pressure-
volume diagrams for propane, propylene, n-butane and iso-butane. Au-
toignition is denoted for n-butane. Data is presented for the Research
method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.6 Representative leaded and unleaded pressure traces for propane and n-
butane at sub-critical compression ratio test conditions, together with the
upper and lower variability limits observed in each batch of 900 cycles. The
amount of TEL added to the intake is reported as a percentage of the total
fuel mass entering the engine. . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.7 Measured engine-out exhaust emissions for CO, NO and total UHCs (wet
basis) versus lambda and compression ratio. Data is presented for the
Research method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . 115
5.8 Combustion efficiency for the twelve fuels with autoignition suppressed
versus lambda. Data is presented for the Research method SKI test condition.117
5.9 The two-zone engine model used in the reverse-run combustion simulation. 118
5.10 GT-Power model of the CFR engine (left to right: intake system, carburettor
venturi, engine cylinder and crankcase, exhaust system). . . . . . . . . . . 120
5.11 Modelling approach used to represent the liquid fuel carburettor in GT-
Power, together with a sectional view of the actual hardware on the CFR
engine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
5.12 Absolute error (% IMEP) between the simulated pressure history and the
representative non-autoigniting pressure trace versus the simulated convec-
tive heat transfer coefficient multiplier. Data is presented for the Research
method SKI test condition, with the cylinder head/piston and cylinder wall
temperatures held constant at 440 K and 410 K respectively. . . . . . . . . 127
5.13 Estimated start (•) and end (◦) of combustion for propane, propylene, n-
butane and iso-butane. Data is presented for the Research method SKI test
condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
5.14 Volumetric efficiency versus compression ratio. Data is presented for the
Research method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . 131
5.15 Measured engine intake temperature under motoring (air only) and firing
(air and fuel) conditions (refer to Figure 3.4b). Data is presented for the
Research method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . 131
5.16 Measured steady-state fuel flow rate versus simulated steady-fuel flow rate.
Data is presented for the Research method SKI test condition. . . . . . . . 132
5.17 Measured engine intake temperature under firing conditions versus simu-
lated cycle-averaged intake temperature under firing conditions. Data is
presented for the Research method SKI test condition. . . . . . . . . . . . . 133
5.18 Measured and reverse-run combustion simulation pressure histories versus
crank angle. Data is presented for the Research method SKI test condition. 134

xxiv
5.18 Measured and reverse-run combustion simulation pressure histories versus
crank angle. Data is presented for the Research method SKI test condition. 135

5.18 Measured and reverse-run combustion simulation pressure histories versus


crank angle. Data is presented for the Research method test condition with
lambda not corresponding to the value for SKI. . . . . . . . . . . . . . . . . 136

5.19 Mass fraction burned histories obtained from the reverse-run combus-
tion simulation without autoignition, together with the location of the
autoignition onset during octane rating. Data is presented for propane (Pr),
propylene (Pryl), n-butane (nB), iso-butane (iB) and mixtures thereof at the
Research method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . 138

5.20 50% mass fraction burned crank angle and 10-90% burn duration. Data is
presented for propane (Pr), propylene (Pryl), n-butane (nB), iso-butane (iB)
and mixtures thereof at the Research method SKI test condition. . . . . . . 139

5.21 Estimated mass fraction burned at the onset of autoignition obtained from
the reverse-run combustion simulation with chemical equilibrium. Data is
presented for propane (Pr), propylene (Pryl), n-butane (nB), iso-butane (iB)
and mixtures thereof at the Research method SKI test condition. . . . . . . 140

5.22 Mass fraction burned at autoignition versus RON. Data is presented for the
Research method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . 140

5.23 Simulated unburned zone temperature versus crank angle for the pure
constituents. Data is presented for the Research method SKI test condition. 141

5.24 Simulated unburned zone temperature versus crank angle for the eight
mixtures. Data is presented for mixtures containing propane (Pr), propylene
(Pryl), n-butane (nB) and iso-butane (iB) at the Research method SKI test
condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

5.25 Experimental autoignition location during octane rating versus simulated


peak unburned zone temperature location. Data is presented for the Re-
search method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . . . 143

6.1 Schematic diagram of the revised two-zone combustion model with detailed
chemical kinetics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

6.2 Flow chart describing the operating principle within the closed portion of
the cycle in the engine model with chemical kinetics. . . . . . . . . . . . . 148

6.3 Estimated residual gas fraction versus compression ratio. Data is presented
for the Research method SKI test condition with autoignition suppressed. 150

6.4 Ratio of the constant-volume ignition delay obtained using the standalone
mechanism to that obtained using the combined mechanism. All simula-
tions were performed with an initial pressure of 40 bar. . . . . . . . . . . . 153

xxv
6.5 Effect of variations in the Woschni convective heat transfer coefficient mul-
tiplier on the simulated autoignition of propane. Positive heat transfer rates
indicate that heat is transferred from the working fluid to the combustion
chamber surfaces. Simulated conditions correspond to the Research method
SKI test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

6.6 Effect of intake air humidity on autoignition. Simulated conditions corre-


spond to the Research method SKI test condition. . . . . . . . . . . . . . . 157

6.7 Simulated variation in the autoignition onset for the pure constituents
with the addition of CO. Filled icons represent the inferred residual CO
concentrations at intake valve closure. Simulated conditions correspond to
the Research method SKI test condition. . . . . . . . . . . . . . . . . . . . . 158

6.8 Engine-out hydrocarbon emissions from a propane-fuelled spark-ignition


engine (adapted from Kaiser et al., 1983). . . . . . . . . . . . . . . . . . . . 160

6.9 Effect of unreacted and partially reacted hydrocarbons on the simulated


autoignition onset of propane. The composition and relative proportions of
these hydrocarbons were modelled using the data reported by Kaiser et al.
(Figure 6.8). Simulated conditions correspond to the Research method SKI
test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

6.10 Simulated constant-volume autoignition delays for the pure constituents,


with varying amounts of NO. All simulations were performed with an initial
pressure of 40 bar, with lambda corresponding to the Research method SKI
test condition for each fuel. . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

6.11 Measured and simulated pressure histories for the pure constituents, with
and without NO. Data is presented for the Research method SKI test condition.165

6.11 Measured and simulated pressure histories for the pure constituents, with
and without NO. Data is presented for the Research method SKI test condition.166

6.12 Simulated variation in the autoignition onset for the pure constituents
with the addition of NO. Simulated conditions correspond to the Research
method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

6.13 Measured and simulated autoignition onset for Propane (Pr), Propylene
(Pryl), n-Butane (nB), iso-Butane (iB) and mixtures thereof. Autoignition
was not observed for propylene when the residual NO concentration was
doubled. Data is presented for the Research method SKI test condition. . . 168

6.14 Simulated fraction of the starting fuel oxidised versus crank angle for
Propane (Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB) and mixtures
thereof. Simulated conditions correspond to the Research method SKI test
condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

xxvi
6.14 Simulated fraction of the starting fuel oxidised versus crank angle for
Propane (Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB) and mixtures
thereof. Simulated conditions correspond to the Research method SKI test
condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
6.15 Simulated species concentration histories for n-butane oxidation versus
crank angle. Simulated conditions correspond to the Research method SKI
test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
6.16 Simulated species concentration histories for iso-butane oxidation versus
crank angle. Simulated conditions correspond to the Research method SKI
test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
6.17 Simulated species concentration histories for propylene oxidation versus
crank angle. Simulated conditions correspond to the Research method SKI
test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
6.18 Simulated autoignition location when the total engine-out NOx emissions
were assumed to be comprised of 95% NO and 5% NO2 . Data is presented
for Propane (Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB) and mix-
tures thereof, at the Research method SKI test condition. . . . . . . . . . . 175
6.19 Simulated fraction of the starting fuel consumed by pre-flame kinetic activ-
ity within the unburned zone versus crank angle. Data is presented for the
Research method SKI test condition. . . . . . . . . . . . . . . . . . . . . . . 177
6.20 Unburned gas temperature increase above the temperature obtained from
the combustion simulation with chemical equilibrium in Chapter 5. Data is
presented for Propane (Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB)
and mixtures thereof, at the Research method SKI test condition. . . . . . . 179
6.20 Unburned gas temperature increase above the temperature obtained from
the combustion simulation with chemical equilibrium in Chapter 5. Data is
presented for Propane (Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB)
and mixtures thereof, at the Research method SKI test condition. . . . . . . 180
6.21 Temperature rise within the unburned zone prior to autoignition arising
from pre-flame kinetic activity versus RON. . . . . . . . . . . . . . . . . . . 181
6.22 Measured and simulated pressure histories for propane under non-standard
octane rating conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
6.23 Measured and simulated pressure histories for propylene under non-standard
octane rating conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
6.24 Relative error in the simulated autoignition onset versus residual NO con-
centration at intake valve closure. . . . . . . . . . . . . . . . . . . . . . . . . 185

B.1 Measured intake and exhaust valve lift profiles for the CFR engine used in
the experimental work. This data was used in the GT-Power gas exchange
sub-models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229

xxvii
B.2 Forward and reverse discharge coefficients for the standard CFR engine
intake and exhaust valves versus the non-dimensionalised valve lift. Data
was acquired using a steady-state flow blench, and was obtained from the
literature (Vancoillie, 2013). . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

D.1 Variation in the total measured exhaust hydrocarbon emissions for propane
as a function of equivalence ratio, along with the corresponding speciated
exhaust hydrocarbon emissions (Kaiser et al., 1983). C2 H6 /C2 H4 = 0.11
(1.0 ≤ φ ≤ 1.3). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
D.2 Total measured exhaust hydrocarbon emissions as a function of compres-
sion ratio. Data was obtained in a CFR engine operated on propane at 1500
rpm, with the spark timing set at 22.0 ◦ BTDC (Kaiser et al., 1984). . . . . . 240

xxviii
List of Tables
1.1 Estimated global technically recoverable crude oil and wet natural gas
resources (United States Energy Information Administration, 2013). . . . . 3

2.1 Estimated LPG vehicle fleet size and number of refuelling sites by region at
year-end 2011 (World LP Gas Association, 2012). . . . . . . . . . . . . . . . 11
2.2 LPG compositional data by region obtained from proprietary MSDSs of BP,
Chevron, ConocoPhillips, ExxonMobil, and Royal Dutch Shell. . . . . . . . 14
2.3 EN 589 and Australian Fuel Determination (Autogas) fuel quality standard
requirements and test methods (European Committee for Standardisation,
2008; Federal Register of Legislative Instruments, 2003). . . . . . . . . . . . 16
2.4 Operating conditions for the Research and Motor test methods (ASTM
International, 2012a,b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5 RON and MON data for the pure constituents from the literature (Puckett,
1945; American Petroleum Institute, 1958; Clifford, 1969; Boldt, 1967). . . . 36
2.6 RON and MON data for LPG mixtures from the literature. Mixture compo-
sitions have been rounded to one decimal place. . . . . . . . . . . . . . . . 36
2.7 MON blend factors specified in the EN 589 (European Committee for Stan-
dardisation, 2008) and ASTM D2598 standards (ASTM International, 2007a). 37
2.8 Summary of kinetic mechanisms in the literature that are relevant to studies
of LPG and conditions at which the mechanisms have been validated. . . 45

3.1 CFR F1/F2 octane rating unit engine specifications (ASTM International,
2012a,b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.2 Summary of dynamometer specifications. . . . . . . . . . . . . . . . . . . . 54
3.3 Operating conditions for the Research and Motor test methods (ASTM
International, 2012a,b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.4 Amount of dilute tetraethyllead added to the gaseous fuels in order to
suppress autoignition at the Research method SKI test condition. . . . . . 69

4.1 Experimentally determined RONs for the pure constituents and equivalent
reported values from the literature (Puckett, 1945; American Petroleum
Institute, 1958; Clifford, 1969). . . . . . . . . . . . . . . . . . . . . . . . . . . 74

xxix
4.2 Experimentally determined MONs for the pure constituents and equivalent
reported values from the literature (Puckett, 1945; American Petroleum
Institute, 1958; Boldt, 1967). . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.3 Two-way Analysis of Variance for the RON and MON data obtained using
the MINITAB statistical package (Minitab Inc., 2010). . . . . . . . . . . . . 80
4.4 Research test method reproducibility limits for 90.0 to 108.0 octane number
(ASTM International, 2012a). . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.5 Predicted MONs for the mixtures studied by Boldt (Boldt, 1967), based on
the regression model in Equation 4.7. . . . . . . . . . . . . . . . . . . . . . . 87
4.6 MON blend factors in the EN 589 (European Committee for Standardisation,
2008) and ASTM D2598 (ASTM International, 2007a) standards, together
with the average measured MONs for the pure constituents obtained in this
study. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.7 Standard error in the prediction of the MON for fuels with different propy-
lene concentrations, using the EN 589 (European Committee for Standardis-
ation, 2008) and ASTM D2598 (ASTM International, 2007a) standards. . . 94
4.8 EN 589 lower vapour pressure limits and corresponding grades (European
Committee for Standardisation, 2008). Limits are specified at a temperature
of 40◦ C. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4.9 Fraction of the various ternary compositional spaces that satisfy the re-
spective requirements of the Australian LPG fuel standard and the EN 589
Grade D fuel specification. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.10 Minimum RONs of fuels that satisfy the compositional or the MON/vapour
pressure requirements of various global LPG fuel standards. . . . . . . . . 100

5.1 Summary of the fuels examined in this chapter, along with the correspond-
ing Research method SKI test conditions. . . . . . . . . . . . . . . . . . . . 106
5.2 Measured engine-out exhaust emissions of CO, total UHCs and NO (wet
basis) for the twelve fuels with autoignition suppressed. Data is presented
for the Research method SKI test condition. . . . . . . . . . . . . . . . . . . 114
5.3 Lower heating values for the pure constituents (Turns, 2000), as used to
calculate the combustion efficiency. . . . . . . . . . . . . . . . . . . . . . . . 116
5.4 CFR engine geometry applied in the GT-Power model. . . . . . . . . . . . 119
5.5 Optimised Woschni convective heat transfer coefficient multiplier (ζ) for
the twelve fuels with autoignition suppressed. . . . . . . . . . . . . . . . . 126

6.1 Estimated residual concentrations of carbon monoxide (CO) and nitric oxide
(NO) at intake valve closure. Data is presented for the Research method
SKI test condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

xxx
6.2 Intake air humidity limits prescribed in Research and Motor methods
(ASTM International, 2012a,b). . . . . . . . . . . . . . . . . . . . . . . . . . 156
6.3 Calculated residual hydrocarbon concentrations at intake valve closure,
based on the data reported for propane (Kaiser et al., 1983). . . . . . . . . . 160
6.4 In-cylinder CO and NO concentrations at intake valve closure for propane
and propylene under standard and non-standard octane rating conditions. 182

A.1 Measured RONs and MONs (primary data set). Mixture mole fractions
have been rounded to one decimal place. . . . . . . . . . . . . . . . . . . . 227
A.2 Measured RONs and MONs (check point data set). Mixture mole fractions
have been rounded to one decimal place. . . . . . . . . . . . . . . . . . . . 228
A.3 Measured RONs and MONs (quaternary mixture data set). Mixture mole
fractions have been rounded to one decimal place. . . . . . . . . . . . . . . 228

B.1 Reference diameters used to non-dimensionalise the valve lift profiles (Van-
coillie, 2013). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230

C.1 Selected values for the water-gas shift reaction equilibrium constant as a
function of temperature (Turns, 2000). . . . . . . . . . . . . . . . . . . . . . 237

xxxi
Chapter 1

Introduction

1.1 The transport sector and climate change

The transport sector has played a central role in the world’s economic development
over the last century. Fuelled by an abundant supply of affordable oil, personal vehicle
ownership has grown to the point where the Australian transport sector now consumes
more than double the energy that it did just three decades ago (Australian Government,
Bureau of Resources and Energy Economics, 2013b). Passenger vehicles have led the
way, accounting for approximately two-thirds of the energy consumed within the sector.
Globally, these trends are expected to continue for the foreseeable future, mainly as the
result of population growth and improved living standards in developing nations (British
Petroleum PLC, 2013; International Energy Agency, 2012; Saito, 2010; Thiel et al., 2010;
Nel and Cooper, 2009).

Transport is still almost entirely dependent upon oil-derived fuels (International En-
ergy Agency, 2012; Diesendorf et al., 2008), and the combustion of these fuels produces
pollutant emissions of varying toxicity. As such, it is estimated that between 50% and
70% of all airborne pollution in urban environments can be attributed to motor vehicles
(Andrews, 2001). However, an increasingly important pollutant originating from the
combustion of carbon-based fuels is carbon dioxide (CO2 ) − a non-toxic substance, but
a prominent greenhouse gas (GHG). Today, there is a broad consensus that GHGs are
contributing to the warming of the earth through the enhancement of the natural green-
house effect. The Intergovernmental Panel on Climate Change (IPCC) has stated that
in order to avoid dangerous levels of climate change, the global mean temperature rise

1
2 Introduction

must be limited to 2.0 ◦ C above pre-industrial levels (Intergovernmental Panel on Climate


Change, 2007). It is estimated that achieving this target would require the atmospheric
CO2 concentration to be stabilised below 450 parts per million (ppm), relative to 400 ppm
today. This would require a considerable reduction in the GHG intensity of all sectors,
including transportation.

1.2 Alternative transport fuels

A substantial increase in the use of alternative transport fuels is considered by some to


be an important part of our response to climate change (Commonwealth Scientific and
Industrial Research Organisation, 2008; Graham et al., 2008). Liquefied Petroleum Gas
(LPG) is one such fuel that can result in lower emissions of regulated pollutants and GHGs
than the more common liquid transport fuels (Oh et al., 2010; Shanmugam et al., 2010;
Mizushima et al., 2009; Pecqueur et al., 2008; Lang et al., 2007; Baker and Watson, 2005;
Watson and Gowdie, 2000; Snelgrove et al., 1996; Klausmeier and Billick, 1993; Hendren,
1983; Menrad et al., 1983; Watson and Milkins, 1982).

LPG is a mixture of variable content, but is primarily comprised of propane and butane.
These hydrocarbons can be liquefied at ambient temperatures, by applying only moderate
pressures. When stored in this form, LPG has an energy density that is comparable to
that of the more common liquid transport fuels, and so has advantages over natural gas
in some applications. In most regions, LPG is also of relatively low cost. It has therefore
become a popular substitute for gasoline and diesel in conventional combustion engines,
with an estimated 17 million vehicles globally capable of operating on LPG at year-end
2011 (World LP Gas Association, 2012).

LPG can be produced during the processing of so-called wet natural gas, and also
during crude oil refining. Future LPG production is therefore dependent on adequate
global reserves of these resources. Emerging contributors to LPG supplies are so-called
unconventional petroleum resources, particularly the shale resources of the United States.
In recent years, major advances in extraction techniques have seen these once near inac-
cessible resources become a significant contributor to US natural gas production (United
1.3 Fuel suitability for spark-ignition engines 3

States Energy Information Administration, 2013). However, few appreciate that these
same resources are also likely to result in the US having an LPG surplus in the coming
decades (United States Energy Information Administration, 2013, 2011a,b; United States
Geological Survey, 2013; British Petroleum PLC, 2013; Braziel, 2012; Gist, 2012; Hart, 2012;
True, 2012). The recently published data presented in Table 1.1 also indicates that these
same resources appear to be abundant globally. It remains to be seen as to whether the
LPG produced from these resources will be used mainly in transportation or other sectors.

Table 1.1: Estimated global technically recoverable crude oil and wet natural gas resources
(United States Energy Information Administration, 2013).

Crude oil Wet natural gas


Resource description (PJ)
(PJ)
Shale/tight oil and shale gas 2.111 × 106 8.540 × 106
Non-shale 1.843 × 107 1.823 × 107
Total 2.054 × 107 2.677 × 107
Increase in total resources due to inclusion of shale oil and gas 11% 47%

Given these considerations, the more widespread use of LPG may also offer opportu-
nities to increase energy security in some regions. Australia, for example, has only limited
domestic supplies of crude oil, and has therefore become increasingly reliant on imported
liquid fuels (Australian Government, Bureau of Resources and Energy Economics, 2013b,a).
However, Australia is a growing net exporter of LPG, owing to increased conventional
natural gas production. Using LPG to displace the more common liquid fuels in such
regions can therefore simultaneously realise both environmental and economic benefits.

1.3 Fuel suitability for spark-ignition engines

It is widely recognised that higher compression ratios increase the fuel economy and
performance of the internal combustion engine. In the case of the spark-ignition engine,
this key design parameter is limited by the increased susceptibility of the engine to
experiencing so-called knock. This undesirable form of combustion is brought about by the
premature autoignition of the unburned air-fuel mixture. Autoignition is an uncontrolled
4 Introduction

form of combustion that results in a rapid release of energy from within the unburned air-
fuel mixture, prior to the arrival of the flame. Often, this phenomenon leads to potentially
damaging rates of in-cylinder pressure rise, and therefore should be avoided.

The autoignition propensity of a spark-ignition engine fuel can be quantified using


the octane number (ON). The octane number is measured using a standardised, variable
compression ratio, Cooperative Fuel Research (CFR) engine (ASTM International, 2012a,b).
The CFR engine is operated at two distinct test conditions, which are used to quantify the
Research and Motor octane numbers (RON and MON respectively) of the fuel. In general,
fuels with higher octane numbers exhibit a greater resistance to autoignition. As such,
these test methods in part establish the fuel’s suitability for use in a spark-ignition engine,
and thus remain central to the production, specification and economic value of gasoline
(Lang et al., 2007; Bradley and Morley, 1997; Heywood, 1988; Wheeler, 1984).

Whilst the octane numbers of gasoline have been studied extensively, far less is known
about the octane numbers of LPG. However, the limited octane data in the literature
shows that some LPGs have higher octane numbers than the more common liquid fuels.
Figure 1.1, for example, indicates that LPGs comprised primarily of propane are likely to
have RONs that are approximately 15 to 20 units higher than standard gasolines. Several
recent studies indicate that the RON is generally most relevant to modern, spark-ignition
engines (Mittal and Heywood, 2009, 2008; Kalghatgi, 2001a,b). As such, the higher RON
of propane permits the use of higher compression ratios, which can result in higher
thermal efficiencies and lower specific GHG emissions. The same benefits are unlikely to
be realised for LPGs comprised primarily of n-butane, owing to its considerably lower
RON. Indeed, variations in the composition of LPG have been known to cause knock in
engines that have been optimised for fuels with significant propane content (Falkiner,
2003; Clifford, 1969).

As discussed later in this thesis, the existing fuel quality standards for LPG generally
do not specify fuel composition. Instead, restrictions are placed on a number of fuel
properties, including the upper/lower vapour pressures and the MON. However, in
contrast with liquid fuels, there is no active test method by which to determine the MON
of LPG, after the Motor (LP) method was withdrawn in 1989 (ASTM International, 1986).
In the absence of a standard test method and a comprehensive set of octane data, the MON
1.3 Fuel suitability for spark-ignition engines 5

Figure 1.1: Research and Motor octane numbers for various pure hydrocarbons, together
with commercial gasolines meeting the requirements of the Australian fuel standards.
Data has been taken from several sources (Federal Register of Legislative Instruments,
2001; Kubesh et al., 1992; Boldt, 1967; American Petroleum Institute, 1958).

of LPG is currently estimated on the basis of fuel composition. This method assumes the
MON is a compositionally weighted linear sum of the MONs of its pure constituents,
but does not appear to have been comprehensively validated for all LPGs. Surprisingly,
there is also no corresponding method that can be used to determine the RON of LPG− a
quantity specified in fuel quality standards for gasoline.

Adding to this lack of octane data, few studies have comprehensively examined the
combustion behaviour of LPG in spark-ignition engines, nor the fundamental mechanisms
responsible for engine autoignition. Whilst several researchers have made contributions in
this area, these studies have generally been confined to engines optimised for gasoline, or
alternatively have only considered a limited range of LPG compositions, e.g. (Mizushima
et al., 2009; Yousufuddin and Mehdi, 2008; Watson and Phuong, 2007; Baker and Watson,
2005; Radu et al., 2005; Campbell et al., 2004; Dutczak and Golec, 2002; Watson and Gowdie,
2000; Gallas, 1959; Moore and Roy, 1959).

Numerical studies of LPG autoignition are even more limited, and have thus far
only considered empirical correlations, e.g. (Chiriac et al., 2006, 2005; Radu et al., 2005).
6 Introduction

Whilst these techniques have proven robust for predicting the onset of autoignition (once
calibrated for specific engine/fuel combinations), they provide no insight into the kinetic
mechanisms responsible for the observed behaviour. Detailed chemical kinetic modelling
of autoignition in spark-ignition engines can address the latter, but thus far does not
appear to have been undertaken for any LPGs.

This thesis therefore intends to make essentially two contributions. First, it will present
a comprehensive study of the RONs and MONs of a range of LPG mixtures, and therefore
address the lack of octane number data that is currently in the literature. It then examines
the physical processes that take place in the CFR engine, and so relates these processes to
the octane numbers of these LPG fuels, with particular attention given to the autoignition
chemistry.

1.4 Thesis layout

This thesis is structured as follows. Chapter 2 discusses the literature that is relevant
to the autoignition and knock of LPG. The various global LPG fuel quality standards
are first summarised. The influence of these standards on the composition and knock
characteristics of LPGs in different regions is emphasised, together with the external
factors that may also influence the composition of LPG. The standard test methods used to
measure the octane numbers of liquid fuels are also discussed, along with the inactive test
method formally used to determine the MON of LPG. The latter is reviewed in relation to
the limited octane data reported in the literature for LPG.

Chapter 2 also discusses the numerical techniques that have been used to study au-
toignition in combustion engines. The advantages and disadvantages of these techniques
are assessed, particularly in relation to multi-zone models that utilise detailed chemical
kinetic mechanisms to simulate autoignition. This includes a discussion of the chemical
kinetic mechanisms in the literature that are relevant to LPG. This information is used to
formulate several research questions for this study.

Chapter 3 describes the equipment and experimental techniques used to acquire the
data presented in the subsequent chapters. The calibration and operation of the CFR
1.4 Thesis layout 7

engine in the standard test methods for octane number is first discussed, along with the
gaseous fuel system that was designed and manufactured. The techniques and software
used to acquire in-cylinder pressure traces and engine-out emissions are then discussed.

A comprehensive set of RON and MON data for LPG mixtures comprised of propane,
propylene, n-butane and iso-butane is presented in Chapter 4. This data is compared with
the limited octane data in the literature, in order to validate the test procedure used in
this study. Statistical techniques are then used to develop, reduce and validate empirical
models that relate LPG composition to its RON and MON. These models are used to
determine the degree of non-linearity, and thus establish whether the linear blending of
octane numbers appears to be reasonable for all LPGs. Finally, the empirical model for the
RON is used to quantify the minimum RONs of LPGs that satisfy the existing global LPG
fuel standards, along with any implications for the design of future LPG fuel standards.

Chapters 5 and 6 then examine how the measured octane numbers of different LPGs
relate to their detailed autoignition chemistry. Chapter 5 presents a full model of the
CFR engine that describes both the open and closed portions of the cycle. The various
sub-models that describe the non-kinetic phenomena are calibrated for each fuel using
in-cylinder pressure data with autoignition suppressed. The key parameters obtained
from these models, namely the flame propagation, the heat losses and the mass that is
estimated to participate in autoignition, are then discussed briefly.

The model of the CFR engine is then revised in Chapter 6 to simulate autoignition
through the inclusion of detailed chemistry in the unburned zone, but with the various
calibrated sub-models of the non-kinetic phenomena retained. This model is used to
systematically examine the effects of heat losses and the different residual combustion
products on pre-flame kinetic activity, the rate of fuel oxidation, along with the autoignition
behaviour of different LPG mixtures during octane rating. Finally, the simulated in-
cylinder pressure histories are compared with those measured at standard test conditions,
where autoignition was observed. This comparison is used to establish which phenomena
need to be included in an accurate model of LPG autoignition in a spark-ignition engine.

The contributions of this work are summarised in Chapter 7. Recommendations are


also made concerning further research that could be undertaken.
Chapter 2

Literature Review

2.1 Introduction

This chapter discusses the literature that is relevant to the autoignition and knock of LPG.
The chapter begins with a brief overview of LPG, including its origins and its use as a
transport fuel. The existing global LPG fuel standards are then discussed, along with
the influence of these standards on the composition and knock characteristics of LPGs in
different regions.

The general problems associated with engine knock and autoignition in spark-ignition
engines are then reviewed, together with the factors that often influence the autoignition
behaviour of common fuels in combustion engines. These are both discussed in relation
to the standard test methods for octane number, and the existing octane data for LPGs.
Finally, the techniques that have been used to study autoignition numerically are examined,
along with the detailed chemical kinetic mechanisms that are relevant to LPG.

2.2 What is Liquefied Petroleum Gas?

Liquefied Petroleum Gas (LPG) is a common fuel in several sectors. It is a mixture of


variable content, but is primarily comprised of propane and butane. When stored in
liquid form, it has an energy density that is comparable to that of the more common
liquid transport fuels, and so has advantages over natural gas in some applications. Like

9
10 Literature Review

natural gas, LPG combustion can also result in lower emissions of regulated pollutants
and greenhouse gases than the more common liquid transport fuels (Oh et al., 2010;
Shanmugam et al., 2010; Mizushima et al., 2009; Pecqueur et al., 2008; Lang et al., 2007;
Baker and Watson, 2005; Snelgrove et al., 1996; Watson and Gowdie, 2000; Klausmeier and
Billick, 1993; Hendren, 1983; Menrad et al., 1983; Watson and Milkins, 1982). LPG is also
usually of relatively low cost. Using LPG to displace gasoline, diesel and kerosene can
therefore simultaneously realise environmental and economic benefits.

2.2.1 The history of LPG

The LPG industry was late to develop in comparison with both the oil and gas industries.
Its origins lie in the Appalachian oil fields of Western Pennsylvania, more than half a
century after oil was discovered in 1859. So-called associated gas had long been produced
in these fields. But without any significant markets for the gas, it was flared or allowed to
weather off.

The discovery of larger oil fields in the early 1900s created an even greater surplus of as-
sociated gas. By this time, domestic markets that were able to consume the gas had started
to develop (Shelley, 2003). However, the distribution of this gas via underground networks
required the gas to be compressed, producing large amounts of so-called condensate. This
condensate was an extremely light gasoline that was referred to as casinghead.

Casinghead was widely available and had a low economic value. It therefore became
a preferred gasoline blending stock for early refiners. However, the lighter fractions in the
casinghead were far too volatile to be used as major components in gasoline, and resulted
in excessive evaporative losses when stored in atmospherically vented tanks (Clifford,
1969; Falkiner, 2003). The lighter fractions in casinghead were later identified as propane,
butane and other hydrocarbons by Walter Snelling of the United States Bureau of Mines.
In 1911, Snelling laid claim to the invention of LPG, and was awarded a patent for his
method of separating the liquid and gaseous constituents in casinghead.
2.2 What is Liquefied Petroleum Gas? 11

2.2.2 The use of LPG as a transport fuel

LPG is already the leading alternative fuel used in the transport sector. In some regions,
its use is so widespread that it is not strictly classified as an alternative fuel. In Turkey,
almost 20% of light duty vehicles operate on LPG, along with more than 2 million vehicles
in each of Korea, Poland and Turkey (Table 2.1). At year-end 2011, it was estimated that in
excess of 17 million vehicles operated on LPG globally, with Australia’s fleet of 650,000
vehicles the sixth largest.

Table 2.1: Estimated LPG vehicle fleet size and number of refuelling sites by region at
year-end 2011 (World LP Gas Association, 2012).

Region Vehicles Refuelling sites


Australia 655,000 3,200
Belgium 50,000 640
Canada 60,000 2,500
China 140,000 310
France 180,000 1,800
India 1,300,000 1,100
Italy 1,700,000 800
Japan 288,000 1,900
Korea 2,300,000 1,600
Mexico 550,000 2,100
The Netherlands 260,000 1,900
Poland 2,230,000 5,900
Russia 600,000 2,000
Thailand 470,000 560
Turkey 2,400,000 8,700
United Kingdom 150,000 1,400
United States 200,000 2,600

Several governments have promoted LPG as a transport fuel as part of their economic
and environmental policy. The Italian government was one of the first. Its market was
developed in the 1950s as an outlet for surplus LPG produced in its domestic refining
industry (Shelley, 2003). Today, Italy’s LPG vehicle fleet has grown to in excess of 1.7
million vehicles. However, the country has also become a net importer of LPG (World LP
Gas Association, 2012). Nonetheless, the Italian government continues to encourage its
use to reduce urban pollution, through a combination of low fuel excise and other fiscal
incentives (World LP Gas Association, 2012; Danielis and Chiabai, 1998). The Republic
12 Literature Review

of Korea similarly promotes LPG for its environmental benefits, but instead through a
combination of reduced fuel excise and direct mandates. Its fleet of around 2.3 million
vehicles is almost exclusively comprised of commercial vehicles.

By contrast, the Turkish and Polish LPG vehicle fleets, the two largest globally, are
dominated by privately owned, low mileage vehicles. Most vehicles have been retro-
fitted− more than 2 million in Poland, and increasing by 250,000 vehicles annually in
Turkey (Radzimirski, 2010). Low vehicle conversion costs remain central to this growth in
both countries, due to the absence of fiscal incentives (Arslan et al., 2010). Similarly, few
incentives are on offer to acquire LPG vehicles in Russia and the United States. In these
markets, the retail price of LPG is comparable to that of both gasoline and diesel (World
LP Gas Association, 2012), making LPG more costly on a calorific basis.

The use of LPG as a transport fuel in Australia is also well established. It was first
encouraged by the government in the 1980s, and therefore the refuelling infrastructure
is well developed. LPG can be purchased at around 3,200 locations nationally, and is
generally less than half the price of both gasoline and diesel (World LP Gas Association,
2012). Vehicle conversion and acquisition grants are also available, with approximately
310,000 provided since 2006 (AusIndustry, 2013).

2.2.3 LPG composition

The composition of LPG varies significantly, and is influenced by several factors. Glob-
ally, approximately 60% of LPG comes from processing so-called conventional natural gas
(Mustovic, 2011). The remainder is mainly produced during the refining of crude oil.
LPG primarily contains paraffins when sourced from natural gas. These include propane
and the butanes, as well as smaller amounts of ethane and longer chain paraffins. LPG
produced during oil refining can also contain olefins, particularly propylene (propene) and
the butylenes (butenes), which originate from various processes. These olefins are some-
times separated and used as feedstocks for other processes (Gary et al., 2007). However,
due to the similar volatilities of propane and propylene, as well as the butanes and the
butylenes, the separation of these constituents is relatively costly and often not undertaken
2.2 What is Liquefied Petroleum Gas? 13

(Sadeghbeigi, 2012; Lamia et al., 2007; Rege and Yang, 2002). As a result, refinery-sourced
LPG can have significant olefin content, with up to 30% by volume reasonably common.

The production of gasoline also influences LPG’s butane content. The economic value
of iso-butane is usually higher as an alkylation feedstock (Gary et al., 2007). This process
produces a very desirable gasoline blending stock. Therefore, generally only surplus
amounts iso-butane are directed into LPG production streams. Similarly, n-butane can be
blended directly into gasoline as a low-cost octane improver, or to regulate the vapour
pressure of the gasoline pool. The availability of n-butane for LPG can therefore vary due
to the seasonally and geographically varying Reid Vapor Pressure (RVP) requirements for
gasoline.

Given these different sources and processing methods, the composition of LPG is
highly variable. Evidence of this can be seen in Table 2.2, which shows the acceptable
range of major species within LPG in different markets around the world, as obtained from
proprietary Material Safety Data Sheets (MSDSs). Whilst LPG content varies significantly
across the globe, it is primarily a mixture of four species: propane, propylene, iso-butane
and n-butane (Figure 2.1). Whilst the ethane content can be up to 10 (% vol) in some
markets, it is a less significant component overall. The butylene content in LPG also
appears to be less significant, but is not widely reported.

2.2.4 LPG fuel quality standards

Variations in the composition of LPG are known to have a significant effect on its knock
characteristics (Falkiner, 2003). However, LPG fuel standards usually do not define com-
positional limits. Instead, restrictions on a number of the fuel properties are generally
specified. European automotive-grade LPG, for example, must comply with the require-
ments of the EN 589 standard (European Committee for Standardisation, 2008). This
standard specifies a minimum Motor octane number (MON) of 89.0, and a minimum fuel
vapour pressure that varies between 275 kPa(g) and 950 kPa(g) throughout the year, to
account for seasonal variations (Table 2.3). The latter requirement is set at a national level,
and therefore may still vary from region to region during a given period. A maximum
14 Literature Review

Table 2.2: LPG compositional data by region obtained from proprietary MSDSs of BP,
Chevron, ConocoPhillips, ExxonMobil, and Royal Dutch Shell.

Fuel composition (% vol)


Region ethane propane propylene butanes butylenes

Australia 0-10 0-100 0-45 0-50 0-15


Belgium 0-5 50-100 0-50 0-10
Canada 92.5-100 0-5 0-2.5
France 0-5 50-100 0-50 0-5
Germany 0-5 50-100 0-50 0-10
Hong Kong 0-5 20-40 60-80
Italy 0-5 40-100 0-50 0-50
Korea 10-35 65-90
Mexico 60 40
Netherlands 0-5 50-100 0-50 0-50
New Zealand 60-70 30-40
Poland 20-60 40-80
Singapore 0-5 20-40 60-80
Spain 0-5 87.5-100 0-5 0-5
United Kingdom 0-5 50-100 0-50 0-5
United States 92.5-100 0-5 0-2.5

fuel vapour pressure of 1550 kPa(g) is also specified as a mechanical limit for fuel storage
and dispensing purposes. Properties such as the evaporation residue and water content
generally relate to vehicle operability considerations, e.g. fuel system corrosion.

The EN 589 standard has also formed the basis of the Australian LPG fuel standard,
but with some differing properties to account for climatic variations (Table 2.3). In contrast
with the European standard, the Australian standard has a fixed lower and upper vapour
pressure limit of 800 kPa(g) and 1530 kPa(g) respectively. These limits do not vary with
season. The Australian standard also has a higher minimum MON requirement of 90.5. In
both standards, these restrictions impose indirect limits on the upper concentrations of
some constituents, but still generally allow significant compositional variations.

In contrast with both the Australian and European fuel standards, the United States
ASTM D1835 standard (ASTM International, 2011) imposes direct compositional restric-
tions on a number of the constituents. This standard requires automotive-grade Special-
duty Propane (equivalent to Propane HD-5 (Gas Processors Association, 1997)) to be com-
prised of at least 92.5 (% vol) propane. Compositional restrictions are also imposed on the
2.3 Knock in spark-ignition engines 15

(a) propane. (b) propylene.

(c) n-butane. (d) iso-butane.

Figure 2.1: Structural formulas for propane, propylene, n-butane and iso-butane.

upper concentrations of propylene and large (C4 +) hydrocarbons of 5 (% vol) and 2.5 (%
vol) respectively. The true Special-duty Propane market for optimised engine applications
has yet to materialise (Falkiner, 2003). Nonetheless, this specification remains the only
LPG fuel widely available in the US.

2.3 Knock in spark-ignition engines

It is well recognised that the highest possible compression ratio maximises the fuel econ-
omy and performance of the internal combustion engine. In the case of the spark-ignition
engine, this key design parameter is limited by the increased susceptibility of the engine
to experiencing so-called knock. This undesirable form of combustion is brought about
by the premature autoignition of the unburned air-fuel mixture, as the result of excessive
pressure and temperature. Often, this leads to potentially damaging rates of in-cylinder
pressure rise, and therefore should be avoided (Heywood, 1988). As a result, considerable
research has been undertaken to overcome knock and explain its underlying mechanisms.
The former has resulted in substantial improvements in fuel quality, whilst also providing
16
Table 2.3: EN 589 and Australian Fuel Determination (Autogas) fuel quality standard requirements and test methods (European
Committee for Standardisation, 2008; Federal Register of Legislative Instruments, 2003).

Property Minimum Maximum Test method


EN 589 (2008)
Motor octane number 89.0 EN 589 Annex B
Vapour pressure (40 ◦ C) a

Grade A kPa(g) 950 1550 EN ISO 8973, EN 589 Annex C


Grade B 800
Grade C 700
Grade D 500
Grade E 275
Total dienes content b % mol 0.5 EN 27941
Total sulfur content c mg/kg 50 EN 24260, ASTM D3246, ASTM D6667
Evaporation residue mg/kg 60 EN 15470, EN 15471
Hydrogen sulphide rating Negative EN ISO 8819
Copper strip corrosion (1 h at 40 ◦ C) rating Class 1 EN ISO 6251
Water content rating Pass EN 15469
Odour rating Distinctive at 20% LFL EN 589 Annex A

Australian Fuel Determination (Autogas) 2003


Motor octane number 90.5 ISO 7941, EN 589 Annex B
Vapour pressure (40 ◦ C) kPa(g) 800 1530 ISO 8973
Total dienes content b % mol 0.3 ISO 7941
Total sulfur content c mg/kg 50 ASTM D2784
Evaporation residue mg/kg 100 JLPGA-S-03
Volatile residues (C5 +) % mol 2.0 ISO 7941
Hydrogen sulphide rating Negative EN ISO 8819
Copper strip corrosion (1 h at 40 ◦ C) rating Class 1 EN ISO 6251
Water content rating Pass EN 15469
Odour rating Distinctive at 20% LFL EN 589 Annex A
a minimum vapour pressure varies by season.
Literature Review

b including 1,3 butadiene.


c after stenching.
2.3 Knock in spark-ignition engines 17

a better understanding of several design techniques that reduce the susceptibility of an


engine to knock. However, the latter has progressed more slowly, and even today, the
mechanisms responsible for knock are still not fully understood.

It is widely accepted that knock arises from a rapid release of energy from within
the unburned air-fuel mixture (so-called end-gas). This region lies ahead of the flame,
and is therefore compressed by both the flame itself, along with the piston. Under some
circumstances, these processes can lead to autoignition. This phenomena is characterised
by the near instantaneous consumption of the end-gas, prior to the arrival of the flame. The
heat release during autoignition occurs up to 25 times more rapidly than during normal
combustion, often causing high frequency pressure oscillations and temperature increases
that exceed 1000 K (Bradley et al., 1994).

2.3.1 Autoignition

The events leading to the autoignition of the end-gas in an optical access spark-ignition
engine are shown in Figure 2.2. Both the flame and any pre-flame kinetic activity are
depicted as dark regions, due to their chemical luminescence. The flame propagates from
left to right, and therefore the end-gas is located to the right of the flame. Pre-flame kinetic
activity is first observed in image 34 at multiple locations within the end-gas. In image 37,
several of these sites have combined to form a single region that exhibits more substantial
kinetic activity. Finally, extensive pre-flame kinetic activity is observed throughout the
end-gas in image 42. The in-cylinder pressure records indicate that this kinetic activity
was accompanied by autoignition.

Experimental and theoretical studies undertaken by Pan et al. (Pan and Sheppard, 1994;
Pan et al., 1998) suggest that three distinct modes of reaction lead to autoignition, namely
deflagration, thermal explosion and developing detonation. Inhomogeneity in thermo-chemical
state of the end-gas plays a central role in determining the prevailing mode, mainly due to
the existence of exothermic centres. These so-called hot-spots exhibit a higher temperature
than their immediate surrounds, and ignite more readily due to the strong temperature
dependence of the chemical reactions (Dreizler and Mass, 1998; Bradley et al., 1994).
18 Literature Review

Figure 2.2: High-speed schlieren cinematography showing the sequence of events leading
to autoignition in an optical access spark-ignition engine (Spicher et al., 1991).

Mixture inhomogeneity has been attributed to several factors, including imperfect mixing,
residual combustion products from previous cycles, along with species that catalyse or
inhibit reactions (Sheppard and Konig, 1990; Bradley and Morley, 1997).

The deflagration mode is relatively benign, and is usually associated with lower
end-gas temperatures. It is characterised by the formation of a small flame front, which
subsequently propagates through the end-gas. In most cases, the deflagration mode
does not result in autoignition (Konig and Sheppard, 1990; Smith et al., 1985). However,
under certain conditions, this mode often transitions into thermal explosion by inducing
additional exothermic centres within the end-gas (Rudloff et al., 2013). Thermal explosion
leads to higher heat release rates than deflagration. Often, this leads to the development of
pressure waves that are characteristic of trace or medium knock intensities (Figure 2.3b).

The developing detonation mode is associated with intermediate temperature gradi-


ents within the end-gas. This mode causes shock waves to develop, and often leads to
damaging rates of in-cylinder pressure rise (Figure 2.3c). As such, developing detonation
is considered the most damaging reaction mode. Prolonged exposure to this mode gener-
ally causes extensive engine damage, including breakage of piston rings, erosion of the
2.3 Knock in spark-ignition engines 19

Figure 2.3: Cylinder pressure histories for (a) normal combustion, (b) medium or trace
knock and (c) heavy knock. Data was acquired in a single-cylinder engine operated at
4000 rpm and wide open throttle (Douaud and Eyzat, 1977).

cylinder head or gasket, and/or melting of the piston (Attard, 2007; Nates, 2000; Fitton
and Nates, 1996; Nates and Yates, 1994; Maly et al., 1990; Heywood, 1988).

2.3.2 Pre-flame kinetic activity

Several early studies demonstrated that pre-flame kinetic activity within the end-gas often
influences the susceptibility of many fuels to autoignite. In one such study, Masson and
Hesselberg used a motored engine to first compress a range of air-fuel mixtures (Mason
and Hesselberg, 1954). These partially reacted mixtures were then recycled into the intake
of a firing engine. The authors observed that an increase in the compression ratio of the
motored engine caused the knock-limited compression ratio (KLCR) of the firing engine
to decrease, thus indicating a higher susceptibility to autoignition. These observations
were consistent with those reported in later studies (Johnson et al., 1965; Moore and Roy,
1959; Walcutt et al., 1954). These studies collectively suggest that autoignition-inducing
compounds are often formed within air-fuel mixtures that undergo compression, as is the
case for the end-gas in a spark-ignition engine.

The heat release associated with pre-flame kinetic activity has also been studied fairly
extensively. Several researchers have used Coherent anti-Stokes Raman spectroscopy
(CARS) measurements to show that these reactions often cause the end-gas temperature
20 Literature Review

to increase considerably for some fuels, e.g. iso-octane, Primary Reference Fuels and
commercial butane (Kalghatgi et al., 1995b; Bradley et al., 1994; Bood et al., 1997). Despite
these often large temperature increases, motored engine studies show that the amount of
fuel consumed in these processes is typically less than 5% of the total charge (Addagarla,
1991; Green et al., 1987a).

2.3.3 Effect of residual combustion products on autoignition

Modern spark-ignition engines often make use of residual combustion products to satisfy
vehicle emissions legislation. The best known example is so-called external exhaust gas
recirculation (EGR). These systems utilise a small pipe and control valve to dilute the fresh
intake air with a small amount of the combustion products from previous cycles. These
gases have a higher heat capacity, and therefore lower the peak combustion temperatures.
This technique is often used to reduce the thermal nitric oxide (NO) production (Galloni
et al., 2012; Turns, 2000; Heywood, 1988). Additionally, residual combustion products
may also remain trapped inside the combustion chamber after closure of the exhaust value
(so-called internal EGR). The internal EGR fraction varies with engine speed, the degree of
throttling and the valve timing (Meyer, 2008)

The products of combustion from a spark-ignition engine are comprised of a range of


chemical species (Heywood, 1988). For hydrocarbon fuels, these include carbon dioxide
(CO2 ), water vapour (H2 O), carbon monoxide (CO), hydrogen (H2 ), along with a range
of unburned hydrocarbons. Smaller amounts of nitric oxide and nitrogen dioxide (NO2 ),
collectively referred to as nitrogen oxides or NOx , may also be present (Lu et al., 2005;
Bhave et al., 2005; Pitts et al., 1984; Nebel and Jackson, 1958). Several studies have shown
that these species often influence both fuel oxidation and autoignition differently, e.g.
(Contino et al., 2013; Brussovansky et al., 1992; Bradley and Yeo, 1993; Affleck et al., 1968).
These effects are summarised as follows.
2.3 Knock in spark-ignition engines 21

Partially reacted hydrocarbons

The post-combustion processes in a spark-ignition engine are characterised by the des-


orption of hydrocarbon-rich gases from various oil layers and fine crevices into which
the flame is unable to propagate (Figure 2.4). Prior to the opening of the exhaust valve,
the combustion chamber temperature is sufficiently high to cause rapid oxidation of the
majority of these hydrocarbons (Heywood, 1988; Heywood et al., 1980). However, once
the blow-down process has commenced, the temperature within the chamber decreases
rapidly. This causes a sudden freezing of the reactions responsible for the oxidation of the
remaining hydrocarbons (Weiss and Keck, 1981).

Due to the variability of these processes, the exhaust gases from a spark-ignition
engine are comprised of many different hydrocarbons (Dober, 2002; Prabhu et al., 1998;
Prabhu, 1997; Kayes and Hochgreb, 1996; Thompson and Wallace, 1994; Finlay et al., 1990;
Heywood, 1988; Tabaczynski et al., 1972). The relative proportions of these hydrocarbons
may also vary considerably with the engine operating conditions (Kaiser et al., 1984,
1983; Ninomiya and Golovoy, 1969; Daniel, 1967). For these reasons, few studies have
systematically examined the influence of partially reacted hydrocarbons on autoignition.
Often, these species have been modelled as methane, or other short-chain hydrocarbons,
e.g. (Dubreuil et al., 2006, 2007). More comprehensive studies have also examined
hydrogen and formaldehyde (Eng et al., 2002; Machrafi et al., 2008). In most cases, these
studies have reported that trace amounts of the abovementioned species generally have an
insignificant influence on the combustion phasing in Homogeneous charge compression
ignition (HCCI) engines.

Nitrogen oxides

The oxidation of many fuels in the presence of NO and NO2 has been studied fairly
extensively. In the case of simple fuels such as hydrogen, carbon monoxide, methane,
ethane, propane and n-butane, the presence of small amounts of NO (<400 ppm) has
been shown to enhance fuel oxidation under most conditions (Rasmussen et al., 2008;
Konnov et al., 2005; Dagaut and Nicolle, 2005; Shin and Yoon, 2003; Dagaut et al., 2001; Tan
22 Literature Review

Figure 2.4: Schematic diagram of the post-combustion processes by which hydrocarbons


desorb and exit the cylinder during (a) the exhaust blowdown process; (b) the exhaust
stroke; (c) the end of the exhaust stroke (Lavoie et al., 1978).

et al., 1999; Nelson and Haynes, 1994; Stein et al., 1992; Bromly et al., 1992; Bromly, 1991;
Ubbelohde et al., 1935; Thompson and Hinshelwood, 1929). This often causes a reduction
in the temperature at which consumption of the fuel commences, with autoignition also
occurring more prematurely than observed for the neat fuels.

The chemical interactions that occur between hydrocarbons and small amounts of NO
are often referred to as a mutual sensitisation. The oxidation of the fuel is enhanced due to
the presence of NO, and similarly the presence of the hydrocarbon further promotes the
conversion of NO to NO2 . This mutual sensitisation has generally been attributed to the
increased production of hydroxyl radicals (OH) that is often observed in the presence of
NO (Dagaut and Nicolle, 2005; Moreac et al., 2006; Hori et al., 1998). These reactive radicals
are produced through the consumption of less active peroxyl radicals (RO2 ), particularly
hydroperoxyl radicals (HO2 , i.e., R = H), via the pathway,
2.3 Knock in spark-ignition engines 23

NO + HO2 → NO2 + OH
(R 2.1)

Reaction R 2.1 indicates that some hydrocarbons will promote the oxidation of NO to NO2
more readily than others. Bromly (Bromly, 1991) observed this behaviour in a flow reactor
as a variation in the so-called cross-over temperature (the temperature at which 50% of the
NO added to the reactor had been oxidised to NO2 ). For example, it was observed that
the cross-over temperature for 68 ppm of propane was lower than that of both 600 ppm
of ethane and 880 ppm of methane. This suggests that a fixed concentration of NO will
more readily oxidise propane than either ethane or methane, under otherwise equivalent
conditions. These observations are consistent with those reported elsewhere for these
hydrocarbons, e.g. Figure 2.5.

Longer chain hydrocarbons tend to interact with NO in a more complex manner than
many simple fuels. For example, several studies have reported that the presence of NO
can either promote or inhibit the oxidation of iso-octane, n-heptane, Primary Reference
Fuels (PRFs), Toluene Reference Fuels (TRFs) and commercial gasolines under certain
conditions (Chan et al., 2012; Anderlohr et al., 2009; Machrafi et al., 2008; Dubreuil et al.,
2006; Burluka et al., 2004; Cairns and Blaxill, 2005; Stenlaas et al., 2003; Stenlass et al., 2002;
Prabhu et al., 1993; Kowalski et al., 1992). Although the prevailing behaviour is strongly
dependent on the temperature, NO concentrations of up to 100 ppm have generally been
observed to enhance fuel oxidation. For NO concentrations of up to 500 ppm, many
researchers have observed that NO begins to have an inhibiting effect on fuel oxidation.
This complex behaviour has also been observed between some olefins and NO, including
1-pentene (Prabhu et al., 1996).

The inhibiting effect of NO on the oxidation of some hydrocarbons can be attributed to


its chain terminating capabilities. The formation of nitrous acid (HONO) via the pathways
described by Reactions R 2.2 and R 2.3, for example, is known to compete with Reaction R
2.1 under certain conditions (Moreac et al., 2006; Stenlass et al., 2002; Bromly, 1991). These
competing pathways strongly influence the hydroxyl radical concentrations, and therefore
24 Literature Review

Figure 2.5: Non-dimensional hydrocarbon concentration versus residence time for


methane (CH4 ), ethane (C2 H6 ), ethylene (C2 H4 ), propane (C3 H8 ) and propylene (C3 H6 )
in the presence of 20 ppm of nitric oxide. Data was acquired in a plug flow reactor at a
temperature of 800 K (Hori et al., 1998).

the availability of these radicals to oxidise the parent hydrocarbon and intermediate
species.

NO + OH (+ M) → HONO (+ M) (R 2.2)

ROOH + NO → RO + HONO
(R 2.3)

Far less is known about the influence of NO2 on autoignition and fuel oxidation. This
is likely due to the small contribution of NO2 to the total engine-out NOx emissions for
spark-ignition engines that operate around the stoichiometric condition (Hilliard and
Wheeler, 1979; Hanson and Egerton, 1937). In what appears to be the most comprehensive
study, Chan et al. observed that NO2 produces similar levels of overall reactivity to NO
(Chan et al., 2011). It was noted, however, that NO tends to enhance the low temperature
2.4 Test methods for the knock characteristics of spark-ignition engine fuels 25

reactivity more significantly, whilst NO2 was more active at higher temperatures. This
behaviour had previously been observed for methane (Bendtsen et al., 2000).

Carbon monoxide

Limited data is reported on the influence of CO on autoignition and fuel oxidation. In


the case of HCCI engines, several studies show that the addition of CO to the intake
system has a negligible influence on both the combustion phasing and the general au-
toignition behaviour (Dubreuil et al., 2006, 2007; Moreac et al., 2006). It is unclear if the
same behaviour occurs in spark-ignition engines, which generally operate at close to
stoichiometric conditions.

2.4 Test methods for the knock characteristics of spark-ignition


engine fuels

The propensity of a spark-ignition engine fuel to knock can be quantified according to its
octane index (OI). This property relates the knocking characteristics of any arbitrary fuel
to a Primary Reference Fuel that exhibits equivalent knocking characteristics at the same
operating condition. The OI is measured in a standard engine at the two fixed operating
conditions, known as the Research and Motor methods. These two OIs are referred to as
the Research and Motor octane numbers (RON and MON) respectively.

2.4.1 The Research and Motor methods

The standard test methods for octane number were developed by the Cooperative Fuel
Research (CFR) committee in the 1920s and 1930s. These test methods utilise a variable
compression ratio, CFR engine (Figure 2.6). These test methods are intended to provide a
link between laboratory-based fuel ratings, and the operating conditions that are likely
to be encountered in a vehicle. The Research and Motor methods were adopted by the
American Society for Testing and Materials (today ASTM International) in 1944 and 1948,
26 Literature Review

Figure 2.6: The Cooperative Fuel Research octane rating unit (Balis, 2013).

under the designations ASTM D2699 and D2700 respectively. Today, these long-standing
test methods continue to have worldwide acceptance, and remain central to the production,
specification and economic value of gasoline (Bradley and Morley, 1997; Wheeler, 1984).

In the Research method, the CFR engine is operated at a speed of 600 rpm (Table 2.4),
with fixed spark timing of 13.0 ◦ BTDC (before top dead centre). The temperature is of the
air entering the carburettor is controlled (52 ◦ C, compensated for variations in barometric
pressure). However, the temperature of the air-fuel mixture downstream of the carburettor
is not designated. For liquid fuels, this causes the temperature of the air-fuel mixture
entering the engine to be lower than that of the air, due to the evaporation of the fuel.
Thus, the Research method is generally considered to correlate most closely with fuel
anti-knock performance under relatively mild operating conditions (ASTM International,
2012a; Bell, 1975).
2.4 Test methods for the knock characteristics of spark-ignition engine fuels 27

Table 2.4: Operating conditions for the Research and Motor test methods (ASTM Interna-
tional, 2012a,b).

Research method Motor method


Operating parameter (ASTM D2699) (ASTM D2700)
Engine speed 600 ± 6 rpm 900 ± 9 rpm
Air intake temperature 52 ± 1.0 ◦ C a 38 ± 2.8 ◦ C
Mixture intake temperature b not specified 149 ± 1.0 ◦ C
Intake air pressure barometric barometric
Intake air humidity 25 - 50 grains H2 O/kg dry air 25 - 50 grains H2 O/kg dry air
Coolant temperature 100 ± 1.5 ◦ C 100 ± 1.5 ◦ C
Oil pressure 172 - 207 kPa(g) 172 - 207 kPa(g)
Oil temperature 57 ± 8.0 ◦ C 57 ± 8.0 ◦ C
Spark timing 13 ◦ BTDC 14-26 ◦ BTDC c
a compensated for variations in barometric pressure.
b refersto the air-fuel mixture temperature measured downstream of the carburettor.
c varied with compression ratio.

In the Motor method, the CFR engine is operated at a higher speed of 900 rpm, with
the spark timing varied between 14-26 ◦ BTDC (altered with octane number). Further
and perhaps most importantly, the temperature of the air-fuel mixture downstream of
the carburettor is regulated to ensure that all fuels enter the engine pre-vapourised, at
a temperature of 149 ◦ C. This requirement removes the thermodynamic charge cooling
for liquid fuels. As a result, the Motor method is generally regarded as a higher severity
test condition, and is assumed to correlate most closely with fuel anti-knock performance
under high speed and load conditions (ASTM International, 2012b; Kumagai and Kudo,
1963)

The difference between the Research and Motor method ratings for a given fuel is
termed the fuel sensitivity, where S = RON − MON. The fuel sensitivity is a measure of the
extent to which a given fuel is downgraded under the more severe operating conditions
associated with the Motor method (Bell, 1975).

Primary Reference Fuels

Primary Reference Fuels (PRFs) are used to define the octane number (ON) scale, as first
proposed by Edgar (Edgar, 1927). These fuels are nominally volumetrically proportioned
28 Literature Review

mixtures of n-heptane and iso-octane (2,2,4-trimethylpentane). These hydrocarbons were


selected as they have comparable boiling points, but exhibit strongly contrasting knocking
characteristics (Bradley and Morley, 1997). Iso-octane exhibits a low knocking tendency,
and is therefore arbitrarily assigned an octane number of 100.0. By contrast, n-heptane
exhibits a high knocking tendency, and is therefore arbitrarily assigned an octane number
of 0.0. A PRF comprised of 70 (% vol) iso-octane and 30 (% vol) n-heptane would have an
octane number of 70.0 (RON and MON).

Many hydrocarbons exhibit superior anti-knock characteristics to iso-octane under


standard octane rating conditions, and therefore have octane numbers that exceed 100.0.
For these fuels, the octane number scale has been extrapolated through the addition of
dilute tetraethyllead (TEL) to iso-octane, using an empirically determined relationship
(ASTM International, 2012a,b). This very effective anti-knock compound, which has
since been outlawed in gasoline in most regions, extends the octane number scale to its
maximum value of 120.3.

2.4.2 The Motor (LP) method

It became apparent in the 1960s that the absence of a standard test method for the knock
characteristics of LPG was causing a great deal of confusion. Earlier publications had
presented a varied picture of the adequacy of LPG as an engine fuel, largely due to the
variations in fuel anti-knock quality throughout the United States (Boldt, 1967; Adams
and Boldt, 1964; Kerley and Thurston, 1956). The Natural Gasoline Processor Association
(NGPA) responded in 1962, producing the composition-based HD-5 fuel specification.
Under this standard, LPG intended for use in spark-ignition engines was required to be
comprised of at least 92.5 (% vol) propane. This standard therefore placed significant
restrictions on fuel composition. Several engine manufacturers and fuel producers consid-
ered the HD-5 standard to be unnecessarily restrictive, and indicated their preference for
a standard based on measured octane number (Boldt, 1967).

In 1964, the ASTM published the preliminary details for a proposed Method of test for
determining knock characteristics of Liquefied Petroleum (LP) Gases by the Motor method. This
2.4 Test methods for the knock characteristics of spark-ignition engine fuels 29

test method was based on the Motor method for liquid fuels, and differed only in the use
of a retro-fitted gaseous fuel carburettor (Figure 2.7). This system was comprised of a
control valve to regulate the fuel flow, along with a flow meter and vaporising unit. The
fuel was first drawn from the storage vessel as a liquid, and vapourised in a water-bath,
that was maintained at a temperature of 82 ± 5.6 ◦ C. The fuel was introduced into the
intake system through an inlet hole on the air horn of the standard liquid fuel carburettor.

Two comprehensive test programs undertaken in 1964 and 1966 culminated in the
ASTM publishing the Motor (LP) method under the designation ASTM D2623 in 1967
(ASTM International, 1986). The test method was subsequently adopted as a joint ASTM-IP
(Institute of Petroleum) standard in 1968, under the designation ASTM D2623/IP 238.

Test method withdrawal

Although the Research and Motor methods for liquid fuels remain active standards to
the present day, the Motor (LP) method was withdrawn by the ASTM in 1989. The test
method continued to be published under the IP designation until 2011, when it was also
withdrawn. Falkiner suggested that its withdrawal could primarily be attributed to the
accuracy by which the MON of LPG could be estimated using linear blending of the
pure component octane numbers (Falkiner, 2003). Additionally, it is also likely that the
compositional limits of the HD-5 fuel standard was sufficiently restrictive that any non-
linear blending effects were too small to be significant. This cannot however be expected
for LPGs of significantly broader composition (Falkiner, 2003), as are presently used in
Australia and globally (Table 2.2).

2.4.3 The Methane number

The Methane number (MN) is a less well known property that has been used to quantify
the knock characteristics of gaseous fuels. This fuel property is also measured using
the standard CFR engine, which is operated in accordance with the Motor method test
conditions (Table 2.4). The test method differs only in the use of methane and hydrogen as
Literature Review
Figure 2.7: Schematic diagram of the gaseous fuel carburettor system used in the Motor (LP) test method (Balis, 2012).

30
2.4 Test methods for the knock characteristics of spark-ignition engine fuels 31

reference fuels, which are arbitrarily assigned MNs of 0.0 and 100.0 respectively. The use
of these reference fuels has been shown to overcome the often inadequate range of the
octane number scale for natural gases (Leiker et al., 1972).

Although not standardised to date, the MN test method offers some benefits for
gaseous fuels. First, the MN scale can be used to include all gaseous fuels (hydrogen,
natural gases, LPGs and their mixtures) without the need for extrapolation (Ryan III et al.,
1993). This practice, as used in the octane number scale above 100.0, tends to make the
test method increasingly insensitive to non-linear fuel effects in the high octane number
range (Falkiner, 2003). Kubesh et al. also showed that the MN can be related linearly to
the MON using Equation 2.1 (Kubesh et al., 1992). However, this relationship does not
appear to have been comprehensively validated for gaseous fuels comprised of 75 (% vol)
methane or less.

MN = 1.624 MON − 119.1 (2.1)

The major shortcoming of the MN scale is its insensitivity to small changes in fuel
composition at the lower end of the scale. Several researchers overcame this shortcoming
by substituting n-butane in the place of hydrogen as the lower reference fuel (Schaub and
Hubbard, 1985; Attar and Karim, 2003). This revision resolved the principal deficiency of
the MN scale for natural gases, but is unlikely to be suitable for all LPGs. Ethane, propane
and n-butane are reported to have MNs of 44.0, 34.0 and 10.0 respectively (Leiker et al.,
1972; Ryan III et al., 1993).

2.4.4 Octane number blending in multi-component fuels

Predicting the octane numbers of multi-component fuels is generally a complicated process.


In most cases, the octane number of a mixture of two or more constituents is not a linear
function of its composition (Bradley and Morley, 1997). As such,

ONmixture 6= X1 ON1 + (1 − X1 ) ON2 (2.2)


32 Literature Review

Since practical fuels are complex mixtures of many hydrocarbons (Knop et al., 2013;
Kalghatgi et al., 2011), the models used to describe the ON blending often reflect only the
interactions between different types of hydrocarbons, e.g. (Iob et al., 1995; van Leeuwen
et al., 1994; Protić-Lovasić et al., 1989). Alternatively, the interactions between various
refinery process streams are often considered, e.g. (Ghosh et al., 2006; Pasadakis et al.,
2006; Albahri, 2003; Albright and Eckert, 1999; Muller, 1992).

In the case of the interactions between different types of hydrocarbons, the blending
of paraffins and olefins generally produces positive deviations from linearity in ON
(Bradley and Morley, 1997). This blending is referred to as synergistic, and is shown in
Figure 2.8. By contrast, the blending of paraffins and aromatics, along with aromatics and
some alcohols, generally produces negative deviations from linearity in ON (Bradley and
Morley, 1997; Lovell et al., 1934; Lovell, 1948). This blending is referred to as antagonistic. In
general, constituents belonging to the same molecular class tend to exhibit linear blending
characteristics in ON (Scott, 1958; Lovell, 1948; Lovell et al., 1934).

Figure 2.8: Linear and non-linear octane number blending between two hydrocarbons, X1
and X2 . Segment (a) indicates synergistic blending, segment (b) indicates linear blending,
and segment (c) indicates antagonistic blending (Ghosh et al., 2006).

Presently, the only models that describe the ON blending of LPGs assume that the
MON is a linear function of the fuel composition. These are discussed in Section 2.5.2.
2.4 Test methods for the knock characteristics of spark-ignition engine fuels 33

2.4.5 The relevance of octane number to modern production engines

The design of the CFR engine differs considerably to modern production engines. Some of
the more antiquated aspects include its flat piston crown, the side-mounted spark-plug,
along with the use of very few turbulence generating mechanisms to enhance combustion.
The engine is also equipped with a fuel carburettor, and operates continuously at wide
open throttle and relatively low speed. These factors cause the engine to knock at a lower
compression ratio than a modern production engine operated on the same fuel (Swarts
et al., 2005).

For these reasons, several studies have examined the relationship between ON and the
performance of various fuels in modern production engines. These studies have tended to
focus on gasolines, with the work of Kalghatgi appearing to be the most comprehensive
(Kalghatgi, 2001a,b). Kalghatgi demonstrated that neither the RON or the MON was able
to adequately describe the fuel anti-knock behaviour for all operating conditions. Rather,
the fuel anti-knock behaviour was best approximated using a combination of the RON
and the MON, or the so-called Octane Index (OI),

OI = (1 − K) RON + K MON = RON − K S, (2.3)

where K is a weighting factor that depends only on the operating conditions. When the
Motor method was introduced in 1932, the average value of K was equal to 1. Therefore,
the value of the OI coincided with the MON. However, there is substantial evidence which
suggests the value of K has steadily decreased since this time (Bell, 1975; Ingamells and
Jones, 1981; Kerley and Thurston, 1956; Boyd, 1938; Sabina, 1938).

Most recently, Mittal and Heywood (Mittal and Heywood, 2009, 2008) performed a
parametric study that considered the influence of various engine operating parameters on
the value of K. These researchers concluded that the value of K would be expected to de-
crease as the air charge entering the engine was increased. These changes are directionally
consistent with modern engine designs, which aim to increase the combustion pressures
to boost efficiency, but lower the combustion temperatures to reduce the susceptibility
to autoignition and knock. This work showed that a modern, naturally-aspirated engine
34 Literature Review

would have an average K value of approximately 0 (Figure 2.9). This outcome produces
an OI that coincides with the RON. Kalghatgi similarly identified that in most cases the
average value of K was slightly negative (Kalghatgi, 2001b). Under such conditions,
Equation 2.3 suggests that the best overall anti-knock performance for a fuel of a given
RON is obtained as the MON is reduced.

Figure 2.9: Estimated variation in the average value of K for spark-ignition engines since
1932 (Mittal and Heywood, 2009).

In the present, smaller capacity, turbocharged, direct injection engines are becoming
increasingly widespread due to their high power density and reduced fuel consumption.
It is expected that these engine configurations are likely to result in a value of K that is
even more negative (Mittal and Heywood, 2009).

2.5 The knock characteristics of Liquefied Petroleum Gas

2.5.1 Research and Motor octane number data

There is very limited octane data for LPG reported in the open literature. Early studies
of octane number that were undertaken between the 1940s and 1950s tended to focus on
gasolines, along with the pure hydrocarbons in the gasoline boiling range. These studies
produced limited and often inconsistent data for the pure hydrocarbons in the LPG boiling
2.5 The knock characteristics of Liquefied Petroleum Gas 35

range (American Petroleum Institute, 1958; Lovell, 1948; Puckett, 1945; Lovell et al., 1934).

A more comprehensive and precise study was undertaken by Boldt (Boldt, 1967). This
work provided an insight into the blending of LPG mixtures and its influence on the
MON. However, most fuels considered in this study contained tetramethyllead (TML),
which has since been outlawed in most regions. The most recent, significant study
appears to be that undertaken for the California Air Resources Board (CARB) in the late
1990s (The Adept Group Inc., 1998, 2000). This work produced octane data for five fuels
comprised of approximately 70-90 (% vol) propane, but lacked detailed compositional
data for the butane isomers, which differ considerably in octane number (only the total C4
concentration was reported). This data is summarised in Tables 2.5 and 2.6.

Table 2.6 indicates that non-linear blending in ON can occur in LPGs. For example, the
addition of 2 (% vol) ethane to propane results in a lower MON than that measured for
pure propane. This suggests that antagonistic blending occurs in binary ethane/propane
mixtures at small ethane concentrations. Additionally, this data also suggests that antago-
nistic blending occurs in binary propane/propylene mixtures. Conversely, the data for
binary propane/n-butane mixtures suggests these constituents blend linearly in ON at
high propane concentrations.

2.5.2 Motor octane number by gas chromatography

In the absence of both an active, standard test method and a comprehensive octane data
set, the MON of LPG is currently estimated from the fuel composition. A single calculation
method has prevailed in both the EN 589 (European Committee for Standardisation, 2008)
and ASTM D2598 standards (ASTM International, 2007a). This methods assumes that the
MON is a compositionally weighted linear sum of the MONs of its pure constituents. For
a sample with q constituents, the MON is estimated as,

q
MON = ∑ Ci Mi , (2.4)
i =1

where Ci is the concentration of the individual constituents in the sample and Mi is the
36 Literature Review

Table 2.5: RON and MON data for the pure constituents from the literature (Puckett, 1945;
American Petroleum Institute, 1958; Clifford, 1969; Boldt, 1967).

Research octane number Motor octane number

Puckett API Clifford Puckett API Boldt


Constituent (1945) (1958) (1969) (1945) (1958) (1967)

ethane 118.5 a 111.7 a - 100.7 a - 99.2± 0.7


propane 112.5 a 111.4 a 110.0 97.1 - 96.3± 0.7
propylene 102.9 a 101.3 a - 84.9 - 83.0± 0.7
n-butane 93.6 94.0 93.5 90.1 89.1 89.0± 0.7
iso-butane 102.1 a 100.3 a 100.4 97.6 - 96.8± 0.7
1-butylene 97.4 97.5 - 81.7 79.9 -
2-butylene (cis) 99.6 100.0 - 86.5 83.5 -
iso-butylene 103.1a - - 88.1 - -

a calculated octane number based on the reported quantity of TEL in iso-octane that produced an
equivalent ASTM Knock Intensity.

Table 2.6: RON and MON data for LPG mixtures from the literature. Mixture compositions
have been rounded to one decimal place.

Mixture composition (% vol) RON MON Source

94.1% propane, 3.7% propylene, 2.2% butane a 108.4 96.1 (The Adept Group, 1998)
85.3% propane, 9.6% propylene, 5.1% butane a 107.7 94.6 (The Adept Group, 1998)
80.2% propane, 14.6% propylene, 5.2% butane a 106.6 93.7 (The Adept Group, 1998)
80% propane, 9.8% propylene, 10.2% butane a 107.0 94.1 (The Adept Group, 1998)
76.1% propane, 3.7% propylene, 20.2% butane a 106.8 94.4 (The Adept Group, 1998)
2% ethane, 98% propane - 95.7 (Boldt, 1967)
2% ethane, 86% propane, 12.0% propylene - 94.2 (Boldt, 1967)
2% ethane, 50% propane, 48% propylene - 89.9 (Boldt, 1967)
95% propane, 5% n-butane - 95.9 (Boldt, 1967)
80% propane, 20% propylene - 92.9 (Boldt, 1967)
4.6% ethane, 93.6% propane, 0.6/1.2% n-/iso-butane 111.0 95.3 (Adams and Boldt, 1964)

a reported butane content is comprised of both n-butane and iso-butane.


2.5 The knock characteristics of Liquefied Petroleum Gas 37

corresponding MON blend factor. In the EN 589 standard, these blend factors are specified
on a mass, volume and molar basis. The latter is used in cases of dispute (European
Committee for Standardisation, 2008). In contrast, ASTM D2598 only allows the MON
to be evaluated on a volume basis (Table 2.7). Given paucity of the MON data in the
open literature for LPG, it is not clear that this calculation method is reasonable for all
LPGs. Further, and quite surprisingly, there is no corresponding method to determine the
RON of LPG, which is specified in fuel quality standards for gasolines (Federal Register of
Legislative Instruments, 2001).

Table 2.7: MON blend factors specified in the EN 589 (European Committee for Standardi-
sation, 2008) and ASTM D2598 standards (ASTM International, 2007a).

EN 589 ASTM D2598

Constituent (mass basis) (mole basis) (volume basis) (volume basis)

ethane 95.9 95.4 95.6 100.7


propane 95.9 95.4 95.6 97.1
propylene 82.9 83.9 83.1 84.9
butane 88.9 89.0 88.9 89.6
iso-butane 97.1 97.2 97.1 97.6
butylenes 76.8 75.8 75.7
pentanes 88.9 89.0 88.9

There are also several limitations associated with the use of these MON calculation
methods. For example, the MON blend factors described in ASTM D2598 are not suitable
for LPGs that contain more than 20 (% vol) propylene (ASTM International, 2007a). Whilst
the EN 589 standard does not specify an equivalent restriction, there is also some doubt
surrounding the validity of this calculation method for LPGs that contain more than 5 (%
vol) propylene and/or butylenes (Australian Government, Department of Environment
and Heritage, 2001; Australian Liquefied Petroleum Gas Association, 2000).
38 Literature Review

Comparison with experimental octane data

The experimental MON data reported by Boldt (Boldt, 1967) is compared with the MONs
evaluated using the EN 589 and ASTM D2598 standards in Figure 2.10. The limited data
indicates that the EN 589 calculation method produces the closest agreement with the
measured octane data, whilst also marginally underestimating the latter. By contrast, the
ASTM D2598 calculation method agrees less well with this experimental data, whilst also
overestimating the latter. Both calculation methods generally yield poor predictions for
the MONs of the pure constituents.

Figure 2.10: Comparison between the MONs computed using the EN 589 and ASTM
D2598 standards, and the experimental octane data reported by Boldt (Boldt, 1967).

2.6 Modelling autoignition in combustion engines

A review of the literature indicates that two primary techniques have been utilised for mod-
elling autoignition in combustion engines. These techniques can broadly be categorised as
either multi-zone or multi-dimensional models.
2.6 Modelling autoignition in combustion engines 39

Multi-zone models are based on classical thermodynamics, and are governed by the
conservation of mass and energy. Mixture homogeneity is therefore assumed within each
distinct zone. Multi-dimensional models, also referred to as fluid dynamic models, are
further governed by the Navier-Stokes equations. These models are characterised by
spatial dependence, with each zone able to exhibit both temperature and compositional
gradients. However, due to the added complexity of the governing equations, it is
computationally very challenging to implement detailed chemistry in multi-dimensional
models (Verhelst and Sheppard, 2009; Bradley and Morley, 1997). Multi-dimensional
models are not considered in the present work. Rather, reference should be made to
several studies in the literature that provide comprehensive treatment of these models
(Liberman et al., 2006, 2005; Natarajan and Bracco, 1984).

2.6.1 Perfectly stirred reactor models with chemical kinetics

The most elementary techniques that have been used to study autoignition generally in-
volve the use of the constant-volume, perfectly stirred reactor (PSR) model, e.g. (Morganti
et al., 2011; Andrae et al., 2005; Chun et al., 1988; Dimpelfeld and Foster, 1984). These
models assume the end-gas is a single zone, which is homogeneous in its thermo-chemical
state. As such, the chemical kinetics are considered the limiting factor within the system.
The PSR model is heated by exothermic reactions, which proceed from a prescribed initial
system state. This raises the pressure and temperature within the reactor, and eventually
leads to the lumped autoignition of the entire mass.

Due to their simplicity, PSR models neglect several processes that occur in real com-
bustion engines. These models are unable to reproduce the compression of the end-gas
by either the flame or the piston. This shortcoming is often addressed through the use
of the so-called variable volume PSR model, e.g. (Iqbal, 2012; Liu, 2010; Dubreuil et al.,
2007, 2006; Faravelli et al., 1998; Bood et al., 1997; Curran et al., 1995; Westbrook and Pitz,
1990; Pitz and Westbrook, 1984). These models operate along similar principles, with the
addition of a prescribed volume history. Thus, this model represents a departure from
the constant-volume PSR model in that the flame propagation is included through its
influence on the pressure history of the end-gas.
40 Literature Review

2.6.2 Multi-zone engine models with chemical kinetics

Multi-zone models are regularly used to perform parametric investigations of new engine
concepts and operating conditions, particularly those susceptible to knock (Verhelst and
Sheppard, 2009). In these models, the end-gas autoignition chemistry is coupled with
both the flame propagation and the motion of the piston. This yields a system in which
the pressure, temperature and species concentrations within the end-gas are temporally
evolving. Further, these models can also be used to describe when autoignition occurs.

General model formulation

The most commonly encountered variant of the multi-zone model is the two-zone model,
which is comprised of a burned and an unburned zone (Figure 2.11). These two zones
are generally separated by an infinitely thin spherical surface of discontinuity, which is
used to represent the flame. This interface is used to transport mass and energy from
the unburned zone to the burned zone, and is generally treated as an adiabatic region in
chemical equilibrium (Liu and Chen, 2009; Hvezda, 2011; Chun et al., 1988). The burned
and unburned zones are characterised by their differing thermo-chemical states. However,
as mentioned previously, each zone is assumed homogeneous in both temperature and
composition.

The pressure throughout the entire control volume is treated as uniform, and is often
prescribed using the pressure histories acquired from engine experiments. This is based
on the assumption that the characteristic time of sound wave propagation across the
combustion chamber is far smaller than the characteristic time of combustion (Moses et al.,
1995). It has been shown that this assumption remains valid until the early stages of the
rapid heat release that generally coincides with autoignition (Jenkin et al., 1997; Witze and
Green, 1993; Konig and Sheppard, 1990).
2.6 Modelling autoignition in combustion engines 41

Figure 2.11: The standard two-zone combustion model framework.

Flame propagation sub-models

Two main approaches have been used to model the flame propagation. In studies where
the in-cylinder pressure history is not available, the flame propagation sub-model is often
prescribed on the basis of turbulent flame speed correlations, e.g. (Eckert et al., 2003;
Jenkin et al., 1997; Hajireza et al., 1998; Chakravarthy et al., 2005). These correlations are
usually based on measured laminar flame speeds, and utilise a series of multipliers to
compensate for variations in the unburned gas temperature and pressure, along with the
influence of diluents on flame speed (Heimel and Weast, 1957; Lancaster et al., 1976).

In studies where the crank-angle resolved in-cylinder pressure history has been ac-
quired, the flame propagation is generally prescribed by fitting this data to a general
functional relationship. The best known example is the so-called Wiebe function (Vibe,
1956), which has been cited extensively throughout the literature, e.g. (Rakopoulos and
Michos, 2008; Swarts et al., 2004; Stenlass et al., 2002; Yates et al., 2003; Dimpelfeld and
Foster, 1984; Leppard, 1985; Hajireza et al., 1998, 1999; Swarts et al., 2005; Soylu, 2005).
This approach is generally favoured, as it is less general and works backwards from a
known result (Verhelst and Sheppard, 2009). The Wiebe function is used to define the
flame propagation based on the fraction of the total fuel mass that has been burned at a
given crank angle,
42 Literature Review

" n + 1 #
θ − θ0

xθ, b = 1 − exp −b
∆θ (2.5)

where xθ, b is the mass fraction burned (MFB) at crank angle θ, ∆θ is the burn duration,
and θ0 is the start of combustion. The constants b and n are tuning parameters that are
adjusted such that the modelled pressure history matches that measured in the engine.

Heat transfer sub-models

The in-cylinder heat transfer that occurs between the working fluid and the combustion
chamber surfaces cannot directly be measured in a standard engine. Therefore, each of the
available techniques that have been used to quantify this parameter must measure the
lumped heat transfer indirectly within the cylinder walls. These techniques often differ
considerably, and have been summarised by several authors (Marr et al., 2010; Demuynck
et al., 2009; Chang et al., 2004; Catania et al., 2000; Guezennec and Hamama, 1999; Borman
and Nishiwaki, 1987).

It is well established that the in-cylinder convective heat transfer coefficient is both
temporally and spatially dependent. However, simple heat transfer correlations that use a
position-averaged method are able to predict the heat transfer between the working fluid
and combustion chamber surfaces in most cases. The most commonly encountered is that
proposed by Woschni (Woschni, 1967). The Woschni correlation without swirl is defined as,

hc = ζ B −0.2 p 0.8 T −0.55 w 0.8 ,


(2.6)

where ζ is the convective heat transfer multiplication factor, B is the cylinder bore, and w
is the mean gas velocity. Once appropriately calibrated, the Woschni correlation has been
2.6 Modelling autoignition in combustion engines 43

shown to be suitable for modelling a wide range of engine and fuel combinations, e.g.
(Baratta et al., 2005; Stenlass et al., 2002; Easley et al., 2001; Catania et al., 2000; Hajireza
et al., 1998; Borman and Nishiwaki, 1987).

Several other correlations for evaluating the convective heat transfer coefficient have
also been proposed, including those by Eichelberg (Eichelberg, 1939), along with Annand
and co-workers (Annand, 1963; Annand and Ma, 1971). However, these correlations
are encountered far less frequently. Comparative studies generally indicate the Woschni
correlation provides the best overall agreement with experimental data (Shayler et al.,
1993; Hayes et al., 1993).

2.6.3 Use of multi-zone models to study autoignition

Multi-zone models developed for the purpose of studying autoignition in combustion


engines are encountered throughout the literature. The two-zone model, for example,
has been utilised to investigate the influence of RON on the autoignition behaviour of
PRFs (Hajireza et al., 1998). This multi-zone model was formulated using empirical
correlations for the heat transfer and the flame propagation. The model was able to
simulate the increasingly prolonged autoignition delay that was observed experimentally
with increasing RON (Figure 2.12).

In a subsequent study, a third boundary layer zone was added to the model (Hajireza
et al., 1999). The addition of the third zone was argued to better predict the thermal
gradients present within the end-gas, and provided good agreement with experimental
CARS temperature measurements. Similarly, Jenkin et al. (Jenkin et al., 1997) accounted
for the near-wall temperature gradients adjacent to the cylinder head, piston crown and
cylinder liner by dividing the end-gas into a series of stacked layers. These zones were
surrounded by an essentially adiabatic core zone.

HCCI engine modelling also provides an important insight into the autoignition
processes in spark-ignition engines. Since the chemical kinetics are the limiting factor,
HCCI engine models often contain eight or more zones, in order to more accurately
predict the end-gas temperature and its spatial distribution. Such modes have been used
44 Literature Review

to successfully predict the autoignition of methane and propane, along with PRFs, TRFs
and natural gas blends, e.g. (Ogink and Golovitchev, 2002; Easley et al., 2001; Fiveland
and Assanis, 2001; Flowers et al., 2000). The major drawback of these models is the added
complexity of the additional zones, which often precludes the use of detailed chemistry.

Figure 2.12: Simulated autoignition onset for PRF40, PRF60 and PRF80 using a two-zone
model with chemical kinetics (Hajireza et al., 1998).

2.6.4 Detailed chemical kinetic mechanisms relevant to LPG

All detailed chemical kinetic mechanisms for natural gases and liquid fuels generally
contain the core oxidation chemistry for C0 -C4 hydrocarbons. But since most of these
models have been developed for larger hydrocarbons, many lack suitable validation for
smaller hydrocarbons, e.g. (Curran et al., 1998, 2002). For these reasons, such models are
often unable to describe the fundamental behaviour of smaller hydrocarbons, including
laminar flame speed and ignition delay times.

Some of the more regularly cited chemical kinetic mechanisms that are relevant to
LPG are summarised in Table 2.8. The most comprehensively validated mechanism in
the open literature appears to be that developed by Healy et al. (Healy et al., 2010a,b,d,c).
This model has been validated using rapid compression machine and shock-tube data for
Table 2.8: Summary of kinetic mechanisms in the literature that are relevant to studies of LPG and conditions at which the
mechanisms have been validated.

Experimental validation conditions


Mechanism Species/reactions Relevant species Temperature (K) Pressure (atm) φ
(Marinov et al., 1998a) C3 H6 , C3 H8a 600-1100 1
(Marinov et al., 1998b) C3 H6 , C3 H8 , n-C4 H10
(Dagaut and Hadj Ali, 2003) 112/827 C3 H6 , C3 H8 , n-/iso-C4 H10 950-1450 b 1b 0.25, 0.5, 1.0, 1.5, 2.0, 4.0 b
(Petrova and Williams, 37/177 C3 H6 , C3 H8 >1000 <100 <3.0
2006)
2.6 Modelling autoignition in combustion engines

(Basevich et al., 2007) 54/288 C3 H6 , C3 H8 , n-C4 H10


(Healy et al., 2010a,b,d,c) 230/1319 C3 H6 , C3 H8 , n-/iso-C4 H10 590-1567 1, 10, 20, 30 0.3, 0.5, 1.0, 2.0
(Naik et al., 2012) 1161/5622 C3 H6 , C3 H8 , n-/iso-C4 H10 c 295-453, 650-1800 1-340 0.3-3.0

a contained a sub-mechanism for NO to NO2 conversion, validated at a concentration of 20 ppm at a hydrocarbon level of 50 ppm.
b validated for a single mixture comprised of 36.2 (% vol) propane, 24.8 (% vol) iso-butane and 39.0 (% vol) n-butane.
c contained a detailed sub-mechanism for NO with NO -formation and fuel-NO sensitisation pathways.
X X X
45
46 Literature Review

propane, n-butane and iso-butane. The laminar flame speeds for these hydrocarbons have
also been validated using data reported elsewhere in the literature. The model developed
by Naik et al. (Naik et al., 2012) has also been validated quite comprehensively, and
contains a sub-model that describes the fuel-NOx sensitisation pathways. However, this
model has not been published in the open literature. Similar models that describe these
same fuel-NOx sensitisation pathways for small hydrocarbons have also been developed
by other researchers, e.g. (Dagaut and Dayma, 2005).

The remaining models in Table 2.8 are generally more limited. Several of these models
do not contain all the species that can be present in LPG, e.g. (Basevich et al., 2007; Petrova
and Williams, 2006; Marinov et al., 1998a,b). Further, several of these models have not
been validated at engine-relevant conditions, e.g. (Dagaut and Hadj Ali, 2003). In the case
of the latter, it is therefore unlikely that this model is suitable for simulating autoignition
in a combustion engine without additional calibration (Westbrook and Pitz, 1990).

To date, the detailed chemical kinetic model developed by Healy and co-workers
has been used to examine the ignition delay of methane/dimethyl ether mixtures, n-
butane/dimethyl ether mixtures, propane/hydrogen mixtures, methane/ethane mixtures,
natural gas/hydrogen mixtures and neat C1 -C4 paraffins (Aul et al., 2013; Brower et al.,
2013; Hu et al., 2013; Man et al., 2013; Zhang et al., 2013; Gersen et al., 2012; Tang et al.,
2012; Beerer and McDonell, 2011). The model has also been used in fundamental studies of
fuel oxidation kinetics and flame speed (Park et al., 2013; Obwald et al., 2011; Lowry et al.,
2011). However, the model does not appear to have been used to examine autoignition in
a combustion engine, nor LPG autoignition.

2.7 Summary and research questions

This chapter reviewed the literature that is relevant to the autoignition and knock of LPG.
Knock is an undesirable form of combustion that occurs in spark-ignition engines. It
is brought about by the premature autoignition of the unburned air-fuel mixture, and
often leads to potentially damaging rates of in-cylinder pressure rise. In general, engine
design and calibration techniques that are employed to increase the efficiency of the
2.7 Summary and research questions 47

spark-ignition engine further increase its susceptibility to knock.

The knock characteristics of liquid fuels are measured with a CFR engine. The octane
number is determined by comparing the knocking tendency of the fuel with that exhibited
by Primary Reference Fuels, at two fixed operating conditions. These tests yield the
Research and Motor octane numbers (RON and MON respectively) of the fuel. There is
no equivalent, active test method that can be used to evaluate the knock characteristics of
LPGs, after the Motor (LP) method was withdrawn in 1989.

Both the Research and Motor octane numbers are generally specified in fuel quality
standards for gasolines. However, in the absence of an active test method for LPGs, the
MON is currently evaluated using a linear sum calculation method, through knowledge of
only the fuel composition. Given paucity of the MON data for LPG in the open literature,
it is not clear that this calculation method is reasonable for all LPGs. Further, there
is no corresponding method that can be used to determine the RON of LPG. This is
surprising, since several studies show that the RON is generally most relevant to modern,
spark-ignition engines.

Autoignition in combustion engines has been studied using a range of numerical


techniques. Multi-zone models coupled with both reduced and detailed chemistry have
been used to simulate the autoignition of methane and propane, natural gases and gaso-
line surrogate fuels. However, no equivalent studies that simulate LPG autoignition in a
spark-ignition engine using detailed chemistry appear to have been undertaken.

Based on this review, the following research questions are proposed.

1. Do the RONs and MONs of propane, propylene, n-butane and iso-butane mixtures
vary synergistically and/or antagonistically?

Based on the literature review, it appears as though no comprehensive studies that ex-
amine the effect of mixture composition on the RON and MON of LPG have been un-
dertaken. The limited experimental data reported in the literature indicates that some
binary propane/propylene mixtures have a lower MON than would be expected based
48 Literature Review

on linear blending behaviour. This suggests that antagonistic blending is possible, which
does not agree with the MON calculation methods specified in the EN 589 and ASTM
D2598 standards. This study will therefore examine whether synergistic and/or antago-
nistic blending occurs in different LPG mixtures, and establish whether these effects are
significant.

2. How do the existing LPG fuel quality standards constrain the RON of LPG in com-
parison with the minimum requirements that are directly mandated in fuel quality
standards for gasolines?

Unlike gasolines, the RON of LPG is not specified in fuel quality standards, despite this
property appearing to be most relevant to modern, spark-ignition engines. The limited
data reported in the literature shows that several of the pure constituents in LPG have con-
siderably higher RONs than all gasolines, whilst others have RONs that are comparable
to standard gasolines. However, it is not clear from this data how the existing MON and
vapour pressure requirements of the Australian and European LPG fuel quality standards
constrain the RON of commercial LPGs in comparison with the RONs that are directly
specified for gasolines. This study will therefore determine the minimum RONs of LPGs
that satisfy the existing Australian and European fuel quality standards, and establish
whether these standards impose unnecessary restrictions on the composition of LPG.

3. What are the phenomena that need to be included in an accurate model of LPG
autoignition in a spark-ignition engine?

The literature review showed that several factors influence autoignition in spark-ignition
engines. Previous studies undertaken for gasoline surrogate fuels indicate that heat losses,
pre-flame kinetic activity and the different chemical species contained in the residual
gases, all influence fuel autoignition differently. However, there does not appear to be
an equivalent study which has examined the influence of these phenomena on LPG
autoignition in a spark-ignition engine. This study will therefore attempt to establish
which phenomena need to be included in an accurate model of LPG autoignition in a
spark-ignition engine.
Chapter 3

Experimental Methods

3.1 Overview

All experimental data presented in this thesis was acquired with a 1933 Waukesha CFR
F1/F2 octane rating unit, located in the Thermodynamics laboratory at the University
of Melbourne. The test facility was comprised of the CFR engine, a 22 kW Siemens
dynamometer, along with other test equipment (Figure 3.1). The CFR engine (1) and
dynamometer (2) were mounted on a concrete plinth (3), that was used to dampen
the mechanical vibrations induced by the test rig. The engine and dynamometer were
primarily operated and monitored from a control bench, located adjacent to the test rig
(4). The gaseous sample fuels were also remotely operated and monitored from this
location, using a controller mounted within the operating panel (5). The operating panel
also housed the standard ASTM octane measurement hardware, including the so-called
Detonation meter (6) and Knockmeter (7) units, as well as a MoTeC Professional Lambda
Meter (PLM) (8). All dynamometer control hardware was housed in a separate electrical
cabinet, and was connected to the rig via an underground access trench (9). The engine
compression ratio (10) and liquid fuel carburettor (11) adjustment mechanisms were
located directly on the CFR engine, and required physical adjustment by the operator.

The standard engine cooling condenser (12) was connected to the building cooling
water system using a standard two-way ball valve. The cooling condenser is convectively-
driven and self-regulating. The boiling of the cooling water is used to maintain the

49
50 Experimental Methods

Figure 3.1: CFR engine test facility layout.

temperature of the engine’s inner cooling surfaces at 100 ◦ C. The oil system was primarily
located internally within the engine. The oil pressure was manually controlled using a
standard two-way ball valve, located on the engine crankcase (13). The oil was heated
using a 240 V crankcase-mounted heater. The oil temperature was also largely self-
regulating once operational.

Air entered the engine via a refrigerated dehumidifier unit supplied by Lawler Corp.
(14). This device replaced the traditional ice-tower unit specified in the standard test
methods, and was used to maintain the air intake humidity within the prescribed limits.
The dehumidifier was connected to the air intake surge tank and heater (15) using a 3 m
section of ribbed aluminium duct, with a nominal diameter of 100 mm (16). The heated air
entered the engine via a vertical pipe, located upstream of the liquid fuel carburettor. This
hardware required some modification to enable the operation of the engine on gaseous
fuels (refer to Section 3.4). The combustion products were expelled from the engine via
a water-cooled exhaust system and surge tank (17) connected to the building exhaust
3.2 CFR Engine 51

extraction system. The surge tank was used to dampen the pressure oscillations induced
by the engine.

In-cylinder pressure data was acquired by removing the standard ASTM Detonation
pickup and installing a Kistler 6125C piezoelectric pressure transducer. The output of the
transducer was preconditioned using a Kistler charge amplifier (18), and acquired using a
LabVIEW data acquisition card, installed in a standard desktop PC. In-house developed
software was used in conjunction with a synchronised Pryde Measurement encoder to
sample the data at intervals of 0.1 crank angle degrees. All data was processed off-line
using MATLAB.

Engine-out exhaust emissions data was acquired with an Auto-diagnostics ADS9000


five gas analyser. This device measured and recorded the concentrations of several
products of combustion. A Bosch LSU 4.9 universal exhaust gas oxygen (UEGO) sensor
was also used to measure the relative air-fuel ratio. This device was powered and operated
using the MoTeC Professional Lambda Meter (PLM).

3.2 CFR Engine

All experimental data presented in this thesis was acquired with a 1933 Waukesha CFR
F1/F2 octane rating unit (Figure 3.2). The CFR engine specifications are listed in Table
3.1. Prior to commencing the experimental work, the engine was fully refurbished and
recommissioned. The cylinder assembly was first removed from the engine, and the liner
was re-bored and honed. The valve guides were also replaced. The engine was rebuilt
using components that comply with the ASTM standards. This included replacement of
the existing aluminium piston with a new cast iron piston, purchased ex-works from the
Waukesha Motor Company in June 2010.

A new engine oil lubrication circuit and water cooling system was also developed. The
latter was connected to the standard condensing unit, and integrated into the building
cooling water system. A new exhaust system was also designed and manufactured. The
design of this hardware dimensionally replicated the exhaust system on the standard CFR
52 Experimental Methods

Figure 3.2: CFR F1/F2 octane rating engine without the gaseous fuel supply system fitted.
3.2 CFR Engine 53

Table 3.1: CFR F1/F2 octane rating unit engine specifications (ASTM International,
2012a,b).

General specifications and geometry


Manufacturer Waukesha Motor Company
Configuration Single cylinder
Crankcase Cast iron, flat combustion chamber roof
Piston Cast iron, four compression rings, single oil control ring
Displacement 37.33 in.3
Bore 3.25 in.
Stroke 4.50 in.
Connecting rod length 10.00 in
Compression ratio Adjustable from 4:1 to 18:1

Valve train
Configuration Single inlet and exhaust valve, push-rod activated
Maximum valve lift 0.238 in.
Intake valve opening location 10 ± 2.5 ◦ ATDC
Intake valve closing location 34 ◦ ABDC
Exhaust valve opening location 40 ◦ BBDC
Exhaust valve closing location 15 ± 2.5 ◦ ATDC

Other
Fuel system Carburettor, single vertical jet (9/16 in. throat)
Ignition system Bosch MEC717 coil and distribution module
Oil SAE-30 monograde
54 Experimental Methods

engine, and was integrated into the building exhaust extraction system. The liquid fuel
carburettor was refurbished with new components, gaskets and alcohol compatible Viton
seals. The engine was also equipped with a new oil heater suitable for a standard 240 V
power supply. The engine was fully re-wired and instrumented with new data acquisition
hardware. Safety guarding surrounding the dynamometer drive-belts, along with the
exhaust system, was also manufactured and installed. An adjustable mounting system
was also manufactured and installed for the dynamometer, which allowed the drive-belts
to be tensioned.

3.3 Dynamometer

The CFR engine was coupled to a Siemens asynchronous motor/generator supplied


by CNC Design, which was used as a dynamometer (Table 3.2). The dynamometer
was controlled and monitored using the proprietary Siemens STARTER software and
a basic operator panel (BOP) located on the control bench, adjacent to the engine. The
STARTER software was installed on a standard desktop PC, and was used to specify the
dynamometer set-point speed. The dynamometer supports a range of control modes, but
was only operated under steady-state conditions in the present work.

Table 3.2: Summary of dynamometer specifications.

Manufacturer Siemens
Model 1PH8137-1MF00-2BA1
Rated power 22 kW
Rated torque 40 Nm
Speed range 0 - 1500 rpm
Maximum torque 350 Nm

The dynamometer was coupled to the CFR engine using three SPC3000 V-belts and
an SPC3 pulley with a nominal pitch diameter of 300 mm. This configuration provided
a speed ratio of 1:1.33, when coupled to the standard flywheel on the CFR engine. The
dynamometer had a rated power and torque of 22 kW and 40 Nm respectively, and could
3.4 Gaseous fuel preparation and delivery system 55

be operated at speeds of up to 1500 rpm. The dynamometer could provide a maximum


torque of 350 Nm, which was sufficient to start the engine. When used as a brake, the
dynamometer operated in regenerative mode and produced electrical energy, which was
exported to the electrical grid. The dynamometer was also capable of motoring the engine.
The dynamometer control system automatically switched between the two modes.

3.4 Gaseous fuel preparation and delivery system

The standard CFR engine is not equipped with the appropriate hardware to measure the
octane numbers of gaseous fuels. Therefore, it was necessary to develop an additional
sample fuel preparation system and modify the standard intake hardware (Figures 3.3-
3.5). The gaseous sample fuel system was comprised of four MKS Instruments 1559A
thermal-based mass flow controllers (MFCs). The MFCs (1) were supplied with factory
calibrations for the individual gases, and were operated by a Type 247D four-channel
controller. Importantly, this controller allowed the overall fuel flow rate to be varied, whilst
maintaining all fuels in fixed proportions. This allowed the maximum knock intensity to
be determined for a given fuel, as per the requirements of the standard test methods for
octane number (refer to Section 3.5).

The gaseous fuels were supplied by Coregas Ltd. in standard D-size and E-size storage
cylinders, and had a minimum purity of 99.5% mol. Each fuel was drawn from the
cylinder as a vapour. Four low pressure regulators (2) located on the cylinders were
used to maintain an upstream pressure of approximately 100 kPa(g). To prevent the less
volatile butanes from condensing within the fuel lines, a closed-loop temperature regulated
heat tracing system was installed. Coalescing filters (3) were also installed immediately
upstream of the butane MFCs to ensure that only vapour entered this hardware. Once the
individual fuels had been metered, the sample fuel was mixed within a 1/2 in. section of
fuel line (4). The gas temperature was measured using a Type T thermocouple inserted
inside the fuel line (5).

The sample fuel was then introduced into the air stream at constant pressure through
a connection on the standard carburettor inlet elbow, upstream of the venturi (Figure
56 Experimental Methods

(a) Gaseous fuel preparation and delivery system.

(b) Gaseous fuel storage cylinders.

Figure 3.3: Gaseous fuel preparation and delivery system used in the experimental work.
3.4 Gaseous fuel preparation and delivery system 57

(a) Isometric view. (b) Sectional view.

(c) Gaseous fuel line entering the engine.

Figure 3.4: Modifications performed to the standard intake system to enable operation on
gaseous fuels.
58 Experimental Methods

3.4). The temperature of the intake air-fuel mixture was measured downstream of the
venturi with a Type T thermocouple. This hardware was inserted into a small hole that
was drilled within the standard ceramic gasket, located between the liquid fuel carburettor
and the engine block (Figure 3.4b). In the event that the test rig emergency stop (E-stop)
system was activated, four solenoid values (6) located in each of the fuel lines disabled the
gaseous fuel supply to the engine.

Overall, this system provided satisfactory engine operation on gaseous fuels. The
system also caused no significant departure from the standard test conditions or hardware
used for measuring the octane numbers of liquid fuels.

3.4.1 Storage cylinder heating

The experimental testing on gaseous fuels was undertaken over a period of approximately
one year. During this period, a wide range of laboratory conditions were encountered.
The ambient temperature, for example, ranged from approximately 10-35 ◦ C. At the lower
end of these temperatures, it was necessary to artificially heat the storage cylinders that
contained n-butane and iso-butane, in order to provide sufficient pressure to operate the
gaseous fuel system. The storage cylinders were positioned in a plastic storage tank with
a water capacity of approximately 30 L (Figure 3.3). Hot water, supplied from a standard
electric water heater, was continuously cycled through the storage tank, and subsequently
drained into the laboratory cooling water system. The artificial heating of the storage
cylinders typically raised the supply pressures to approximately 300-400 kPa(g). This
system also prevented the storage cylinders from cooling as fuel was drawn off to operate
the engine.

3.4.2 Leak detection system

A gas leak detection system was also installed within the test facility to detect hydrocarbon
vapours. The system was comprised of three sensors supplied by Neodym Technologies.
Each sensor was calibrated for individual gases, and placed in strategic locations around
3.4 Gaseous fuel preparation and delivery system

Figure 3.5: Piping and instrumentation diagram for the gaseous fuel preparation and delivery system.
59
60 Experimental Methods

the test facility. The first sensor was positioned at ground level. A second sensor was
positioned in the service trench, and a third sensor was positioned above the engine for
gases that were lighter than air. The sensors each contained an integrated normally open
(NO) relay latch. The latches were wired into the test rig’s electrical E-Stop system. In the
event that any of the sensors detected hydrocarbon vapours, the solenoid valves located
in each of the four gaseous fuel supply lines were closed.

3.4.3 Fuel flow measurement calibration verification

A Micro Motion ELITE CMF010M Coriolis flow meter was used to validate the calibration
of the four MFCs prior to commencing the experimental work. The same procedure
was repeated at the conclusion of the experimental work. The Coriolis flow meter was
positioned in series with the MFCs (Figure 3.6) and measured the mass flow rate of each
gas, ideally independent of the fuel and its thermodynamic state. The data was acquired
using a LabVIEW analogue input channel, by converting the 4−20 mA output on the
transmitter to a 0.996−4.98 V signal using a high precision 249.0 Ω resistor (± 1%).

Figure 3.6: The Coriolis flow meter used to verify the calibration of the MFCs at the
commencement and completion of the experimental work.
3.5 Standard test methods for octane number 61

Prior to validating the calibration, each MFC was operated for a period of one hour
to ensure that all critical operating parameter were stable (MKS Instruments, 1996). The
fuel system was then purged, re-zeroed and pressured to 100 kPa(g). The engine was
subsequently operated under motoring conditions, and simultaneous steady-state fuel
flow measurements were acquired from the two devices over the full range of condi-
tions relevant to the experimental work (Figure 3.7). This data indicates that very good
agreement was obtained both before and after the experimental work was undertaken.

3.5 Standard test methods for octane number

The Research and Motor test methods (ASTM International, 2012a,b) quantify the knocking
tendency of spark-ignition engine fuels at two fixed operating conditions (Table 3.3). In
both test methods, the octane number of an arbitrary fuel (referred to herein as the
sample fuel) is determined by comparing its knocking tendency with that of two Primary
Reference Fuels (PRFs), which are of a defined octane number (refer to Section 3.5.2).
To induce knocking combustion on the sample fuel, the compression ratio of the engine
is gradually increased. This causes a transition from smooth engine operation, to an
operating condition where autoignition of the end-gas occurs. The compression ratio
is non-linearly linked to the octane number of the fuel via the Guide Tables (ASTM
International, 2012a,b). Therefore, a fuel that produced a given Knock Intensity (KI) at a
higher compression ratio will generally exhibit a higher octane number.

The sample fuel, along with the two PRFs, are each rated at the air-fuel ratio that pro-
duces the maximum KI value for each fuel. This property is measured on the proprietary
ASTM Knockmeter in arbitrary divisions of 0 to 100. The Knockmeter is connected to a
proprietary signal conditioning system and so-called Detonation pickup, that is mounted
within the combustion chamber. The air-fuel ratio that produces the maximum KI value
for each fuel is termed Standard Knock Intensity (SKI). Locating the SKI condition in-
volves sweeping the air-fuel ratio, by raising and lowering the carburettor fuel bowls, and
observing the peak value on the ASTM Knockmeter. When two PRFs have been identified
that produce a higher and lower KI value than the sample fuel, the octane number of the
62 Experimental Methods

(a) propane. (b) propylene.

(c) n-butane. (d) iso-butane.

Figure 3.7: Measured fuel flow rates using the MFCs versus the Coriolis flow meter for
the four pure constituents.
3.5 Standard test methods for octane number 63

sample fuel can be determined by interpolation,

KI PRFL − KI gaseous fuel


 
ON gaseous fuel = ON PRFL + (ON PRFH − ON PRFL )
KI PRFL − KI PRFH (3.1)

where ON sample fuel , ON PRFL and ON PRFH are the octane numbers of the sample fuel, the
PRF with the lower octane number, and the PRF with the higher octane number respec-
tively. Similarly, KI gaseous fuel , KI PRFL and KI PRFH represent the equilibrium Knockmeter
values obtained for the sample fuel, the PRF with the lower octane number, and the PRF
with the higher octane number respectively. This test procedure is undertaken twice (with
the PRFs rated in the reverse order), and an average estimate of the octane number of the
sample fuel is made. The test is declared to be a valid octane number rating if a number of
conditions are satisfied. These are specified in the ASTM standards (ASTM International,
2012a,b).

3.5.1 Engine Fit-for-Use procedure

Prior to rating the sample fuels, engine compliance was established in accordance with
the standard ‘Fit-for-Use’ procedure using Toluene Standardisation Fuels (TSFs) of known
accepted reference value (ASTM International, 2012a,b). This procedure was undertaken
at the start of each operating period, or when the engine had been shut-down for a period
of more than two hours. Air intake temperature tuning was not required at any stage
during the experimental work to obtain octane ratings within the specified limits of the
standard test methods.

3.5.2 Liquid reference and standardisation fuel preparation

PRFs and TSFs were prepared on a gravimetric basis using a precision balance with a reso-
lution of 0.001 g. The mass equivalents of the volumetrically-defined blend components
64 Experimental Methods

Table 3.3: Operating conditions for the Research and Motor test methods (ASTM Interna-
tional, 2012a,b).

Research method Motor method


Operating parameter (ASTM D2699) (ASTM D2700)
Engine speed 600 ± 6 rpm 900 ± 9 rpm
Air intake temperature 52 ± 1.0◦ C a 38 ± 2.8 ◦ C
Mixture intake temperature b not specified 149 ± 1.0 ◦ C
Intake air pressure barometric barometric
Intake air humidity 25 − 50 grains H2 O/kg dry air 25 − 50 grains H2 O/kg dry air
Coolant temperature 100 ± 1.5 ◦ C 100 ± 1.5 ◦ C
Oil pressure 172 − 207 kPa(g) 172 − 207 kPa(g)
Oil temperature 57 ± 8.0 ◦ C 57 ± 8.0 ◦ C
Spark timing 13 ◦ BTDC 14 − 26 ◦ BTDC c
a compensated for variations in barometric pressure.
b temperature measured downstream of the carburettor.
c varied with compression ratio.

were determined based on the density of the individual components at 15.56 ◦ C (60.0 ◦ F)
(ASTM International, 2012a,b). Iso-octane (ASTM reference fuel grade) was supplied by
Haltermann GmbH. PRFs with octane numbers that exceeded 100 were prepared using
dilute tetraethyllead (TEL) Type G fuel anti-knock compound supplied by Innospec Inc.
The dilute TEL mixture contained 18.23 ± 0.05% mass tetraethyllead (ASTM International,
2012a,b). The balance of the mixture was comprised of ethylene dibromide (chemical
scavenger), xylene and n-heptane (diluents) and small quantities of various dyes and
insert compounds. All other chemicals used in the experiments were analytical grade and
had a minimum purity of 99.0% mol.

3.5.3 Modified test method for gaseous fuels

In contrast with liquid fuels, there are currently no active test methods for measuring
the octane numbers of LPG. Therefore, the octane numbers of the gaseous fuels were
determined using a method that was consistent with the Equilibrium Fuel Level Bracketing
procedure described in the standard test methods for liquid fuels (ASTM International,
2012b,a). This modified test method is consistent with that specified in the Motor (LP)
test method (ASTM D2623) for LPG, which was withdrawn in 1989 (ASTM International,
3.5 Standard test methods for octane number 65

1986).

The modified test method was used to measure both the Research and Motor octane
numbers of all LPGs mixtures. The test method utilised the standard liquid fuel carburettor
system, along with two PRFs that bracketed the knocking tendency of the gaseous sample
fuel. When the engine was operated on the gaseous sample fuel, the selector valve on the
liquid fuel carburettor was moved into the ‘off’ position. This disabled the flow of liquid
fuel to the engine. Similarly, the gaseous fuel system was disabled when operating the
engine on the PRFs and TSFs. The test procedure is summarised as follows:

1. The engine was thoroughly warmed to ensure that all operating parameters were stable and within
the prescribed limits specified in Table 3.3.

2. The engine ’Fit-for-Use’ procedure was performed using a TSF applicable to the octane number range
where the sample fuel was expected to rate.

3. Engine calibration was undertaken to establish SKI using a PRF with an octane number expected to be
close to that of the sample fuel. The guide tables were then used to adjust the cylinder height to the
barometrically compensated dial indicator value corresponding to the octane number of the calibration
PRF.

4. The air-fuel ratio was varied in order to determine the value for the SKI for the calibration PRF by
adjusting the fuel bowl height. The METER dial on the Detonation meter was then adjusted to produce
a reading of 50 ± 2 divisions on the Knockmeter.

5. The engine was subsequently operated on the gaseous sample fuel. The air-fuel ratio was varied in
order to determine the value for the SKI. The cylinder height was then adjusted to produce a reading
of 50 ± 2 divisions on the Knockmeter. The equilibrium Knockmeter reading was recorded.

6. The octane number of the sample fuel was estimated using the guide tables, based on the dial indicator
setting compensated for the prevailing barometric pressure. The first PRF (PRF1 ) was then prepared
by selecting an octane number that is close to the estimated value of the gaseous sample fuel. The
engine was then operated on PRF1 . The air-fuel ratio was varied in order to determine the value for
SKI by adjusting the fuel bowl height. The equilibrium Knockmeter reading was recorded.

7. A second PRF (PRF2 ) that was expected to produce a Knockmeter reading that bracketed the Knockmeter
reading of the gaseous sample fuel was then prepared. The engine was operated on PRF2 . The air-fuel
ratio was varied in order to determine the value for SKI by adjusting the fuel bowl height. If the
Knockmeter reading of the two PRFs bracket that of the sample fuel, the equilibrium Knockmeter
reading was recorded. Otherwise, additional PRFs were used until this requirement was satisfied.
66 Experimental Methods

8. The same tests were performed to obtain repeat Knockmeter readings for: (1) the gaseous sample fuel,
(2) PRF2 and (3) PRF1 (in the order specified).

9. The octane number of the gaseous sample fuel was calculated using Equation 3.1 for each set of
Knockmeter readings. The mean value of the two calculated octane numbers was assumed to constitute
a valid octane rating provided: (1) the difference between the two calculated octane numbers is not
greater than 0.3, (2) the mean value of the two Knockmeter readings for the sample fuel is between 45
and 55, and (3) the PRFs were within the appropriate octane number limits of the sample fuel specified
in the standard test methods.

10. The dial indicator reading was checked to ensure the value (compensated for the prevailing barometric
pressure) was within the prescribed limits of the guide table value for the calculated octane number
of the gaseous sample fuel. This condition was satisfied when the dial indicator reading was within
±0.014 in. of the corresponding guide table value.

3.6 In-cylinder pressure measurement

In-cylinder pressure data was acquired using a Kistler 6125C vibration compensated and
heat shielded piezoelectric pressure transducer. The pressure transducer was mounted
within a custom machined, flush mounted adapter. The adapter was mounted in the hole
that normally accommodates the standard ASTM Detonation pickup (Figure 3.8). The
pressure transducer has an operating range of 0−300 bar and -20−350 ◦ C. Therefore, to
enable measurements above the maximum operating temperature, the pressure transducer
was thermally insulated by applying a 0.5 mm layer of high temperature silicon paste
onto the sensing diaphragm.

The output of the transducer was preconditioned using a Kistler Type 5007 charge
amplifier, and acquired using a LabVIEW data acquisition card installed in a standard
desktop PC. The transducer was statically calibrated using a dead-weight tester. In-house
developed software (Dennis and van Deventer, 2008) was used in conjunction with a
synchronised Pryde Measurement encoder to sample the data at intervals of 0.1 crank
angle degrees. This software was also used to reference the transducer output to an
absolute pressure measurement acquired in the intake manifold (so-called pressure pegging
(Davis and Patterson, 2006)). The encoder was mounted externally to the engine crankshaft
3.6 In-cylinder pressure measurement 67

Figure 3.8: Kistler 6125C piezoelectric pressure transducer used to acquire in-cylinder
pressure data.

using a custom-machined mild steel adapter with an internal press fit. The run-out in the
crankshaft was accounted for within the encoder mounting, which allowed small amounts
of vertical translation to occur where necessary.

3.6.1 Acquiring in-cylinder pressure data with autoignition suppressed

To estimate the crank angle resolved mass fraction burned (MFB) history, it was necessary
to acquire in-cylinder pressure data where the charge was fully consumed by turbulent
flame propagation in the absence of autoignition (refer to Chapter 5). This prerequisite
was satisfied by preparing a fuel mixture that contained a small amount of dilute TEL.

The TEL was added to a 1.0 mL burette (0.1 mL/graduation) mounted on the standard
air intake pipe. It was introduced into the air stream through a connection, using an 18G
(1.2 mm x 40 mm) hypodermic needle (Figure 3.9). The needle was located approximately
200 mm upstream of the gaseous fuel inlet to ensure that the TEL was thoroughly mixed
with the air prior to the introduction of the gaseous fuel. A small needle valve that was
68 Experimental Methods

Figure 3.9: Hardware used to add dilute TEL to the gaseous fuels.

connected to the burette using 1/4 in. flexible tube was used to adjust the quantity of TEL
added to the intake.

When the desired engine operating condition had been reached, the flow rate of the
TEL was gradually increased until autoignition was no longer observed on the data
acquisition computer used to acquire the in-cylinder pressure data. At this time, the
amount of TEL in the burette was recorded, and the engine was operated under steady-
state conditions for a period of approximately three to five minutes. The quantity of TEL
consumed from the burette during this period was used to estimate the average flow rate
using a stop-watch. Between adjacent test cases, the engine was operated on iso-octane
for approximately five minutes to ensure that any carry-over TEL residue did not affect
the subsequent test.

The amount of dilute TEL added to each gaseous fuel in order to fully suppress
autoignition is summarised in Table 3.4. It is evident that pure propylene, as well as fuels
with significant propylene content, each required larger amounts of dilute TEL to suppress
autoignition. This is consistent with data reported in the literature, which indicates that
lead-tetra alkyls are less effective at suppressing autoignition for olefins (Downs et al., 1951;
3.7 Measurement of the engine-out exhaust emissions 69

Bradley and Morley, 1997). The effect of TEL on the in-cylinder pressure data acquired for
propane and n-butane is presented in Figure 3.10. Similar characteristics were observed
for the other fuels examined in this work.

Table 3.4: Amount of dilute tetraethyllead added to the gaseous fuels in order to suppress
autoignition at the Research method SKI test condition.

Fuel composition (% mol)


dilute TEL
propane propylene n-butane iso-butane (% mass) b
100.0 - - - 0.75
- 100.0 - - 2.67
- - 100.0 - 0.30
- - - 100.0 0.37
70.0 10.0 10.0 10.0 1.02
66.6 16.7 - 16.7 0.99
66.6 - 16.7 16.7 0.83
50.0 50.0 - - 1.59
50.0 - 50.0 - 0.71
41.7 16.6 41.7 - 0.51
16.7 - 66.7 16.7 0.55
- - 50.0 50.0 0.36
b percentage of the total fuel mass entering the engine.

3.7 Measurement of the engine-out exhaust emissions

The dry engine-out exhaust concentrations of oxygen (O2 ), carbon monoxide (CO), carbon
dioxide (CO2 ), nitric oxide (NO) and the total unburned hydrocarbons (UHCs) were
measured using an Auto diagnostics ADS 9000P five gas analyser. This device uses
the Nondispersive Infrared Detector (NDIR) principle to detect CO, CO2 and UHCs,
whilst an electro-chemical cell method is used to detect O2 and NO. The HC concentrations
reported using NDIR techniques are generally not very precise, due to the wide variety of
hydrocarbons that may be present within the exhaust. Each of these hydrocarbons posses
different light absorption characteristics, and therefore it is common practice to correct
these measurements using a multiplier that is dependent upon the fuel (Andrianov, 2011;
Dober, 2002; Heywood, 1988). This was undertaken in this study.
Experimental Methods
Figure 3.10: Measured in-cylinder pressure data for propane and n-butane at the Research method SKI test condition, with and
without TEL. Data is presented for 300 consecutive cycles.

70
3.8 Measurement of the relative air-fuel ratio 71

This exhaust gas analyser continuously sampled the concentrations of the aforemen-
tioned species. The sampling line was connected to the exhaust sampling port via a 1/4 in.
flexible tube. Before entering the gas analyser, the exhaust gas sample was dried using an
in-line diesel fuel filter that acted as a water trap. The hardware was calibrated in-house,
and utilised in-house developed software to acquire and store the data. A comprehen-
sive discussion of the operating principles of the device and the calibration procedure is
provided by Andrianov, 2011.

3.8 Measurement of the relative air-fuel ratio

The relative air-fuel ratio (λ) was measured using a Bosch LSU 4.9 wideband universal
exhaust gas oxygen (UEGO) sensor. This device has an operating range of λ=0.65 to
ideally infinity (Bosch GmbH, 2011), and is therefore suitable for the full range of operating
conditions encountered in this work (λ: 0.90-1.03). The operating principles of the sensor
are discussed elsewhere in the literature (Andrianov, 2011; Regitz and Collings, 2008).

The UEGO sensor was installed within the exhaust system, upstream of the surge
tank. A MoTeC Professional Lambda Meter (PLM) was used to power and operate the
sensor, whilst also providing the sensor output via an inbuilt screen. Due to the catalytic-
based operation of the UEGO sensor, this device was positioned downstream of the
sampling port for the Auto-diagnostics ADS 9000P five gas analyser discussed previously.
The sensor was calibrated at regular intervals using the so-called ‘free air’ calibration
procedure (Bosch GmbH, 2011).
Chapter 4

The Research and Motor Octane


Numbers of Liquefied Petroleum Gas

4.1 Introduction

This chapter presents an experimental study of the Research (RON) and Motor (MON)
octane numbers of Liquefied Petroleum Gas (LPG). A comprehensive set of RON and
MON data for mixtures of propane, propylene, n-butane and iso-butane are presented,
using a method that is consistent with the currently active ASTM Research and Motor
test methods for liquid fuels. Empirical models which relate LPG composition to its
RON and MON are then developed, such that the simplest relationships between the
constituent species’ mole fractions and the mixture octane rating are achieved. This is
used to determine the degree of non-linearity between the composition and the RON and
MON of different LPG mixtures. Finally, implications for LPG fuel quality standards are
discussed briefly, as part of a suggested, more substantial undertaking by the community
which also revisits the standard test procedures for measuring the RON and MON of LPG.

4.2 Pure constituents

The RONs and MONs for propane, propylene, n-butane and iso-butane were first mea-
sured. For each constituent, an average RON and MON was determined based on eight

73
74 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

measurements obtained over approximately a six month period. These values are reported
in Tables 4.1 and 4.2 respectively, together with the limited data reported in the literature.

Table 4.1: Experimentally determined RONs for the pure constituents and equivalent
reported values from the literature (Puckett, 1945; American Petroleum Institute, 1958;
Clifford, 1969).

Research octane number


This study Puckett API Clifford
Constituent (1945) (1958) (1969)
propane 109.4 112.5 111.4 110.0
propylene 100.2 102.9 101.3 -
n-butane 93.5 93.6 94.0 93.5
iso-butane 100.1 102.1 100.3 100.4

Table 4.2: Experimentally determined MONs for the pure constituents and equivalent
reported values from the literature (Puckett, 1945; American Petroleum Institute, 1958;
Boldt, 1967).

Motor octane number


This study Puckett API Boldt
Constituent (1945) (1958) (1967)
propane 96.3 97.1 - 96.3± 0.7
propylene 83.3 84.9 - 83.0± 0.7
n-butane 89.0 90.1 89.1 89.0± 0.7
iso-butane 96.8 97.6 - 96.8± 0.7

In general, good agreement is observed between this study and the literature. The best
overall agreement is with the more recent data of Clifford (Clifford, 1969) and Boldt (Boldt,
1967) respectively, where each of the discrepancies are within the reproducibility limits
of the standard tests at the respective octane levels (ASTM International, 2012a,b). The
average MONs obtained for each constituent are also within the 95% confidence limits of
Boldt (Boldt, 1967), which were established using data obtained from twelve independent
laboratories. Each of the four pure constituents have RONs and MONs that exceed the
minimum requirements for unleaded gasoline in Australia of 91.0 and 81.0 respectively
(Federal Register of Legislative Instruments, 2001).
4.3 Binary mixtures 75

4.3 Binary mixtures

The RONs and MONs for the six binary combinations of propane, propylene, n-butane
and iso-butane are presented in Figure 4.1. The experimental data shows good agreement
with the limited mixture data reported by Boldt (Boldt, 1967). Two of these mixtures
contained 2 (% vol) ethane (Figure 4.1a), but exhibit no significant deviation from the
experimental trends for the binary propane/propylene mixtures.

The blending characteristics are shown to depend primarily on the molecular class of
the constituents. The mixtures comprised of only paraffins (propane, n-butane, iso-butane)
generally exhibit linear blending characteristics under both Research and Motor method
conditions. However, the mixtures comprised of a paraffin blended with propylene exhibit
greater non-linearity, which approaches approximately 1.5 octane numbers in some cases
(e.g. Figures 4.1a and 4.1e).

4.4 An octane number model for LPG

The functional relationship between the composition of LPG (X1 , ..., Xq ) and octane
number is presently unknown. In most cases, the octane number of a mixture of fuels is
not a linear function of its composition (Bradley and Morley, 1997). As such,

ONmixture 6= X1 ON1 + (1 − X1 ) ON2 . (4.1)

Polynomials are regularly used as approximating functions, with a wide variety of


models examined in the literature. The most commonly encountered appear to be Scheffé’s
polynomials (Becker, 1968; Dimitrov and Kamenski, 1999; Draper and Pukelsheim, 1998;
Scheffé, 1958). These models are particularly flexible, and are often adapted to accommo-
date the specific blending characteristics in a given system,
76 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

Figure 4.1: Experimentally determined RONs and MONs for binary mixtures of propane,
propylene, n-butane and iso-butane together with data reported by Boldt (Boldt, 1967).
4.4 An octane number model for LPG 77

q q q
η = ∑ β i Xi + ∑ ∑ βij Xi Xj + ∑ ∑ ∑ βijk Xi Xj Xk
i =1 i< j i < j<k (4.2)
+ ... + β 12 ... N X1 X2 ... X N .

Higher-order terms are often added to describe more elaborate characteristics (e.g. syner-
gistic/antagonistic blending, non-symmetry, minima/maxima offset from the centroid,
etc). These terms are often necessary in mixture models, particularly when the experimen-
tal region spans the entire compositional space (Montgomery, 2012).

4.4.1 Experimental design and statistical analysis

The RONs and MONs for up to ternary mixtures of propane, propylene, n-butane and iso-
butane were studied by three-level factorial experiments. Each constituent in the mixture
is represented by its mole fraction, Xi , where X1 , X2 , X3 , X4 are propane, propylene,
n-butane and iso-butane respectively. The mole fractions are not independent variables,
as they are required to sum to unity,

q
0 ≤ Xi ≤ 1, ∑ Xi = 1.
i =1 (4.3)

The mixtures were prepared on a molar basis using a so-called augmented {3,2}
simplex lattice design (Cornell, 2002), as presented in Figure 4.2. The simplex contained
a total of ten design points for each ternary mixture system. Four of these design points
were located within the interior of the simplex, and represented ternary mixtures. The
vertices and edges of the simplex represented the pure constituents and binary mixtures
respectively.

The RONs and MONs at these design points were measured using randomised experi-
mental runs. This technique generally ensures that the assumption of error terms which
78 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

Figure 4.2: The augmented {3,2} simplex lattice design used for the mixtures in this study.
Mixtures are presented on a molar basis.

are independently distributed random variables is valid (Montgomery, 2012). This condi-
tion is a requirement of the statistical methods used throughout the model development
procedure. The raw octane number data is presented in Appendix A.

4.4.2 Regression analysis and response surface fitting

Regression analysis was performed to determine the partial regression coefficients (β) and
the significance of model terms. The 26 design points within the four ternary factor spaces
were fitted to models of increasing complexity using the MINITAB statistical software
package (Minitab Inc., 2010). The adequacy of the fitted regression models was evaluated
using a range of model summary statistics, including the adjusted coefficient of determi-
nation (R2adj ) and the mean square error (MSE ). Lack of fit tests and an examination of the
error terms for standard regression assumptions (Montgomery, 2012) were used as further
indicators of model adequacy. The latter included consideration of the reproducibility
limits for octane number in the standard test methods (ASTM International, 2012a,b),
which were used as an arbitrary indicator of excessive variability in the error terms.
4.4 An octane number model for LPG 79

Based on these considerations, Scheffé’s quadratic and special quartic polynomials


were identified as the most appropriate models for approximating the functional relation-
ship that existed between the composition and the octane numbers. The special quartic
model (Equation 4.4) was fitted to the RON data, whilst the lower-order quadratic model
(Equation 4.5) was fitted to the MON data. In the limiting case of a fuel that contained
only a single constituent, the regression analysis was performed such that these models
reproduced the pure constituent octane numbers previously presented in Tables 4.1 and
4.2.

q q q
η = ∑ β i Xi + ∑ ∑ βij Xi Xj + ∑ ∑ ∑ βijk Xi2 Xj Xk
i =1 i< j i < j<k
q q
(4.4)
+∑∑∑ β ijk Xi X 2j Xk + ∑∑∑ β ijk Xi X j Xk2 .
i < j<k i < j<k

q q
η = ∑ β i Xi + ∑ ∑ βij Xi Xj .
i =1 i< j (4.5)

A two-way Analysis of Variance (ANOVA) was used to study the effects of the various
intermolecular interactions on octane number, and reduce the fitted regression models
where possible. Tests for equality of variance (Bottenberg and Ward, 1963) were used to
compare the fitted regression models with a series of reduced models, where individual
or groups of predictor variables had been omitted. Statistical hypothesis testing was
then undertaken using the p-value statistic. The latter is defined as the smallest level of
statistical significance at which the data are significant (Montgomery, 2012). A threshold
value of 0.05 was chosen as an acceptable probability of a Type I error (Chow, 1997).
Several statistically insignificant (p>0.05) terms were identified in the fitted regression
models and were eliminated.

The ANOVA demonstrated that a larger number of predictor variables were required
for the RON data to provide insignificant lack of fit. The reduced regression model
80 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

Table 4.3: Two-way Analysis of Variance for the RON and MON data obtained using the
MINITAB statistical package (Minitab Inc., 2010).

Source of Degrees of Adjusted sum Adjusted mean F p


variation freedom of squares square

RON
Fitted with special quartic model
Regression 9 431.37 47.93 956.89 <0.001 a
Residual 16 0.780 0.049 >0.05
Total 25 432.15

MON
Fitted with quadratic model
Regression 6 440.83 73.47 1937.58 <0.001 a
Residual 19 0.752 0.038 >0.05
Total 25 441.58

a statistically highly significant.

for the RON, as presented in Equation 4.6, contained a total of ten terms. Six of these
terms described the non-additive blending behaviour that was present in these mixtures,
including four binary terms and two quaternary terms. Two of the binary terms indicate
antagonistic blending. The remaining terms indicate synergistic blending. However, the
quaternary effects terms are only significant in propane/propylene/butane mixtures at
high propylene concentrations. The remaining four linear terms describe the effect of the
respective pure constituents on octane number.

RON = 109.4 X1 + 100.2 X2 + 93.5 X3 + 100.1 X4 − 5.55X1 X2 − 4.31 X1 X3

+ 2.64 X2 X3 + 4.94 X2 X4 + 59.48 X1 X22 X3 + 44.15 X1 X22 X4 . (4.6)

The reduced regression model for the MON is presented in Equation 4.7. By contrast,
only three non-additive blending terms were required to provided insignificant lack of fit.
Each of the non-linear predictor variables is a binary term with a magnitude greater than
zero. This indicates that only synergism occurs in these mixtures. The fewer number of
4.4 An octane number model for LPG 81

interaction terms in Equation 4.7 indicates that the MON is more linear than the RON for
these mixtures.

MON = 96.3 X1 + 83.3 X2 + 89.0 X3 + 96.8 X4 + 2.79 X1 X4 + 3.53 X2 X3

+ 6.26 X2 X4 . (4.7)

The overall adequacy of the reduced models is considered in Figure 4.3. These plots
demonstrate that both the RON and the MON for the 26 design points are well predicted
by the respective models. The standard error (SE), which is defined as,

v
N
u

u 2
SE = t ONi, measured − ONi, predicted / N,
i =1 (4.8)

is similar for each test condition (RON, SE=0.15 and MON, SE=0.16). Further, the largest
deviations from the experimental measurements are approximately equal for the two test
methods, and do not exceed 0.4 octane number.

4.4.3 Model validation

Model predictive capability for new observations

Confirmation experiments were performed to demonstrate that new observations were


correctly predicted by the regression models. The models were deemed adequate if they
satisfied two criteria. First, the standard error in the model predictions of octane number
should demonstrate no significant deviation from the observed value for the primary
data set used to develop the models. Second, the error terms should have a mean of
approximately zero, with a variance that is independent of the predicted octane number.
82 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

(a) Research octane number. (b) Motor octane number.

Figure 4.3: Measured octane number versus predicted octane number for the design point
data, based on the regression models in Equations 4.6 and 4.7.

An additional set of ‘check point’ observations (Cornell, 2002) were collected for 24
distinct mixtures using randomised experimental runs. These mixtures were comprised
of two and three constituents, with each differing in composition to the design points
used to develop the models (Figure 4.4). The regression models were used to predict the
octane numbers for these check points (Appendix A). These values are compared with the
experimental measurements in Figure 4.5.

Overall, the models demonstrate good predictive capability for these mixtures. There
is no evidence of any significant outliers at either test condition that could indicate model
inadequacy, nor the presence of excessive error terms that correspond to particular octane
levels. The standard errors for each test condition (RON, SE=0.13 and MON, SE=0.14) also
correspond closely with those observed for the primary data set (Figure 4.3).

Model predictive capability for quaternary mixtures

The predictive capability of the regression models was further assessed on eight sample
fuels that contained each of the four constituents. These quaternary mixtures had two
fixed compositional proportions. The first contained a primary constituent that accounted
4.4 An octane number model for LPG 83

Figure 4.4: Design and check point locations in ternary factor space. Mixtures are presented
on a molar basis.

(a) Research octane number. (b) Motor octane number.

Figure 4.5: Measured octane number versus predicted octane number for the check point
data, based on the regression models in Equations 4.6 and 4.7.
84 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

for 70 (% mol) of the mixture, while the second contained a primary constituent that
accounted for 50 (% mol) of the mixture. The balance of each mixture was comprised of
equal proportions of the three remaining constituents. The raw octane number data for
these mixtures is presented in Appendix A.

The regression models were used to predict the octane numbers for these quaternary
mixtures (Figure 4.6). The predicted octane numbers for each mixture correlates well
with the experimental measurements. Further, the standard errors for each test condition
(RON, SE=0.14 and MON, SE=0.15) do not deviate significantly from the corresponding
values for the primary data set used to develop the models. This is despite the absence
of quaternary mixture data in the regression analysis. This suggests that the blending
characteristics exhibited within the quaternary mixtures do not differ considerably to
those in mixtures with up to three constituents.

(a) Research octane number (b) Motor octane number

Figure 4.6: Measured octane number versus predicted octane number for quaternary
mixtures, based on the regression models in Equations 4.6 and 4.7.

Model compliance with the standard test method reproducibility limits

The experimental reproducibility limits were used as a further indicator of model adequacy.
These limits quantify the expected variability in octane number during the compliant
4.4 An octane number model for LPG 85

operation of the Research and Motor test methods (ASTM International, 2012a,b). The
reproducibility limits differ in magnitude between the two test methods, and vary with
octane number. The reproducibility data for the Research method spans the range 90.0 to
108.0 octane number (Table 4.4). The data for the Motor method is more limited, and does
not cover all octane levels relevant to this study. The available data covers the range 80.0
to 90.0 octane number, for which a reproducibility limit of 0.7 octane number is specified.

Table 4.4: Research test method reproducibility limits for 90.0 to 108.0 octane number
(ASTM International, 2012a).

RON Reproducibility
limit (ON)
90.0 - 100.0 0.7
101.0 1.0
102.0 1.4
103.0 1.7
104.0 2.0
104.0 - 108.0 3.5

Analogous to the definition of these limits in the standard test methods, an adequate
model was deemed to be one in which the maximum variability in the error terms did not
exceed the reproducibility limits at each octane level. The error term (residual) for the ith
observation is the difference between the measured and the predicted octane numbers,

residual = ONi, measured − ONi, predicted .


(4.9)

The error terms for the primary and check point data are presented in Figure 4.7. The
error terms for the RON and the MON data each lie within the reproducibility limits of
the standard test methods. Importantly, there is also no apparent bias in error distribution,
nor a correlation when plotted in order of acquisition.
86 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

(a) Research octane number. (b) Motor octane number.

Figure 4.7: Predicted octane number versus error terms for all experimental data acquired
in this study, based on the regression models in Equations 4.6 and 4.7. The reproducibility
limits for the standard test methods are provided where available (ASTM International,
2012a,b).

Model predictive capability for octane data in the literature

The predictive capability of the MON regression model was subsequently assessed using
octane data in the literature. The model was used to predict the MONs of four mixtures
studied by Boldt (Boldt, 1967). Two of the mixtures contained small amounts of ethane,
which was ignored in this analysis. These mixtures were approximated as binary mixtures
with equivalent propane-to-propylene proportions as the original mixtures. The mixture
compositions were then converted from the reported volume basis to a molar basis. The
predicted MONs for these mixtures are presented in Table 4.5.

The model predictions for these mixtures generally agree well with the experimental
data reported by Boldt. For the majority of fuels, the predicted MONs deviate from
the experimental data by 0.4 octane number or less. This includes the two mixtures
that contained ethane. A single fuel comprised of 80 (% vol) propane and 20 (% vol)
propylene lies marginally outside the 95% confidence limits reported by Boldt. However,
the observed discrepancy of 0.6 octane number is still within the reproducibility limits of
4.5 RONs and MONs for up to ternary mixtures 87

Table 4.5: Predicted MONs for the mixtures studied by Boldt (Boldt, 1967), based on the
regression model in Equation 4.7.

Motor octane number

Regression Boldt
Fuel composition (% vol) model (1967)

2% ethane, 86% propane, 12% propylene a 94.6 94.2± 0.4


80% propane, 20% propylene 93.5 92.9± 0.4
2% ethane, 50% propane, 48% propylene a 89.7 89.9± 1.2
95% propane, 5% n-butane 96.0 95.9± 0.5
a fuels were approximated as binary propane/propylene mixtures.

the Motor method at this octane level (ASTM International, 2012b).

4.5 RONs and MONs for up to ternary mixtures

The regression models for the RON and the MON are presented as ternary contour dia-
grams in Figures 4.8 and 4.9 respectively. These contour diagrams exhibit varying degrees
of curvature, which indicates the octane numbers of these mixtures is generally non-linear
with respect to composition. The presence of propylene in these mixtures is the major
source of non-linearity in the RON. However, the effect of this constituent is noticeably
different when blended with a combination of propane/butane or exclusively with bu-
tanes. This can be attributed to the different blending characteristics exhibited between
propylene and the various paraffins. As previously shown in Figure 4.1, propylene blends
antagonistically with propane, but synergistically with both n-butane and iso-butane.

Figures 4.8a and 4.8b demonstrate that these competing effects give rise to complex
curvature in the compositional space approaching pure propylene. In contrast, Figure
4.8c demonstrates that the single ternary mixture which contained only paraffins exhibits
almost linear blending over the entire compositional space. This behaviour is consistent
with a number of studies in the literature, which indicate that constituents belonging
to the same molecular class generally exhibit linear blending characteristics in octane
88 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

(a) propane/propylene/n-butane. (b) propane/propylene/iso-butane.

(c) propane/n-butane/iso-butane. (d) propylene/n-butane/iso-butane.

Figure 4.8: Contour diagrams of RON for ternary mixtures of propane, propylene, n-
butane and iso-butane generated using the regression model in Equation 4.6. Data is
presented on a molar basis.
4.5 RONs and MONs for up to ternary mixtures 89

(a) propane/propylene/n-butane. (b) propane/propylene/iso-butane.

(c) propane/n-butane/iso-butane. (d) propylene/n-butane/iso-butane.

Figure 4.9: Contour diagrams of MON for ternary mixtures of propane, propylene, n-
butane and iso-butane generated using the regression model in Equation 4.7. Data is
presented on a molar basis.
90 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

number (Ghosh et al., 2006; Lovell et al., 1934; Lovell, 1948). The MONs of these mixtures
are generally more linear than the RONs. In the majority of compositional spaces, this
non-linearity generally does not exceed an octane number.

4.5.1 Fuel sensitivity

The sensitivity of a fuel is defined as the difference between its RON and MON,

S = RON − MON. (4.10)

Substituting Equations 4.6 and 4.7 into Equation 4.10, the fuel sensitivity can therefore be
expressed as,

S = 13.1 X1 + 16.9 X2 + 4.5 X3 + 3.3 X4 − 5.55X1 X2 − 4.31 X1 X3

− 2.79 X1 X4 − 0.89 X2 X3 − 1.32 X2 X4 + 59.48 X1 X22 X3 (4.11)

+ 44.15 X1 X22 X4 .

The fuel sensitivity exhibited by the ternary mixtures is presented in Figure 4.10. The
most significant contributors to fuel sensitivity are propane and propylene (S=13.1 and
S=16.9 respectively). The butanes, however, exhibit considerably less fuel sensitivity
(n-butane, S=4.5 and iso-butane, S=3.3). The five binary effect terms similarly contribute to
reduced fuel sensitivity. However, these terms are generally most significant in mixtures
that already exhibit large sensitivities (e.g. propane/propylene mixtures, S=5.55). The two
quaternary effect terms contribute towards increased fuel sensitivity, but are only signifi-
cant in ternary propane/propylene/butane mixtures at high propylene concentrations.
4.5 RONs and MONs for up to ternary mixtures 91

(a) propane/propylene/n-butane. (b) propane/propylene/iso-butane.

(c) propane/n-butane/iso-butane. (d) propylene/n-butane/iso-butane.

Figure 4.10: Contour diagrams of fuel sensitivity for ternary mixtures of propane, propy-
lene, n-butane and iso-butane generated using the regression model in Equation 4.11. Data
is presented on a molar basis.
92 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

4.6 Validity of the EN 589 and ASTM D2598 MON calculation


methods

The EN 589 (European Committee for Standardisation, 2008) and ASTM D2598 (ASTM
International, 2007a) standards specify the same method for estimating the MON of LPG
through knowledge of its composition only. This method involves calculation of an octane
number for each constituent based on fixed blend factors. Therefore, this method assumes
the MON is a linear function of composition. For a sample with q constituents, the MON
is estimated as,

q
MON = ∑ Ci Mi ,
i =1 (4.12)

where Ci is the concentration of the individual constituents in the sample and Mi is


the corresponding MON blend factor. In the EN 589 standard, these blend factors are
specified on a mass, molar or volume basis. In cases of dispute, the molar factors are
used to determine the MON (European Committee for Standardisation, 2008). By contrast,
ASTM D2598 only allows the MON to be evaluated on a volume basis. If required,
interconversion of the sample composition is performed in accordance with ASTM D2421
(ASTM International, 2007b).

The values of Mi relevant to this study are reported in Table 4.6. The blend factors
generally differ between the two standards (e.g. propane and propylene by 1.5 and 1.8
octane numbers respectively). A number of constituents are also assigned blend factors
which differ considerably to the measured MONs (e.g. propylene by up to 1.6 octane
numbers).

The MONs determined using the EN 589 and ASTM D2598 standards are compared
with the measured values in Figure 4.11. For each calculation method, the standard
error exceeds half an octane number. The blend factors specified in the ASTM D2598
standard consistently over-predict the MON. Binary propane/propylene mixtures are the
4.6 Validity of the EN 589 and ASTM D2598 MON calculation methods 93

Table 4.6: MON blend factors in the EN 589 (European Committee for Standardisation,
2008) and ASTM D2598 (ASTM International, 2007a) standards, together with the average
measured MONs for the pure constituents obtained in this study.

Motor octane number/blend factor


ASTM This
Constituent EN 589 a EN 589 b
D2598 b study
propane 95.4 95.6 97.1 96.3
propylene 83.9 83.1 84.9 83.3
n-butane 89.0 88.9 89.6 89.0
iso-butane 97.2 97.1 97.6 96.8
a molar basis.
b volume basis.

largest source of error, with the disparity for these mixtures ranging between 1.0 and 1.6
octane numbers. By contrast, the blend factors specified in the EN 589 standard generally
under-predict the MON. This appears to be, in part, due to the low MON blend factor
assigned to propane, relative to the measured value. The average disparity between the
two calculation methods is approximately 1.1 octane numbers.

(a) Motor octane number by EN 589. (b) Motor octane number by ASTM D2598.

Figure 4.11: Measured MON versus predicted MON obtained using the EN 589 (European
Committee for Standardisation, 2008) and ASTM D2598 (ASTM International, 2007a)
calculation methods.
94 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

4.6.1 Limitations of the MON calculation methods

It is specified in the ASTM D2598 standard that the MON calculation procedure is ap-
plicable only to LPG that contains 20 (% vol) propylene or less. There is no equivalent
restriction specified in the EN 589 standard. However, there is some doubt surrounding
the validity of the MON blend factors for fuels with high olefin (propylene and butylenes)
concentrations (Australian Government, Department of Environment and Heritage, 2001;
Australian Liquefied Petroleum Gas Association, 2000).

The adequacy of the EN 589 and ASTM D2598 standards for estimating the MON of
fuels that contain propylene is examined in Table 4.7. For the fuels considered in this
study, the EN 589 standard error has an equivalent value (SE=0.65) on either side of the
20 (% vol) propylene threshold. This value also shows no significant departure from the
standard error for all fuels (SE=0.63). Therefore, there is no evidence that suggests the
EN 589 MON calculation method produces higher levels of error for fuels that contain
propylene concentrations exceeding 20 (% vol), relative to other fuels.

Table 4.7: Standard error in the prediction of the MON for fuels with different propylene
concentrations, using the EN 589 (European Committee for Standardisation, 2008) and
ASTM D2598 (ASTM International, 2007a) standards.

Motor octane number SE


ASTM
Fuel composition (% vol) N EN 589
D2598
≤ 20% propylene a 15 0.65 0.54
> 20% propylene 24 0.65 0.85
All fuels studied 58 0.63 0.70
a fuels that contained non-zero propylene content.
4.7 RONs of LPG satisfying existing fuel standards 95

4.7 RONs of LPG satisfying existing fuel standards

4.7.1 Australian Fuel Determination (Autogas)

The Australian LPG fuel standard has fixed vapour pressure limits of 800 kPa(g) and
1530 kPa(g) (specified at 40 ◦ C) that do not vary with season, along with a minimum
MON of 90.5 (Federal Register of Legislative Instruments, 2003). Compliance with these
requirements for a given fuel is determined in accordance with ISO 8973 (International
Organisation for Standardisation, 1997) and ISO 7941/EN 589 (International Organisation
for Standardisation, 1988; European Committee for Standardisation, 2008) respectively.
The RONs for ternary mixtures of propane, propylene, n-butane and iso-butane that satisfy
these requirements are presented in Figure 4.12.

The vapour pressure and MON requirements impose significant restrictions on the
compositional space. However, these restrictions still do not entirely eliminate the non-
linearity that can be present in the RON. Further, the two requirements generally influence
the composition in different ways. The minimum vapour pressure has the most significant
influence on the maximum butane content in compliant fuels, whilst the minimum MON
requirement is generally the most influential factor in mixtures that contained propylene.
The upper vapour pressure limit does not influence the mixture compositions considered
in this study, due to the absence of minor constituents that exhibit considerably higher
volatilities than the primary constituents (e.g. ethane and ethylene). The minimum RON
of any fuel that complies with the 90.5 MON and 800 kPa(g) requirements is 99.5 (Figures
4.12a and 4.12d). This value is significantly higher than the minimum requirement of 91.0
for unleaded gasoline in Australia (Federal Register of Legislative Instruments, 2001).

4.7.2 European EN 589 standard

The European EN 589 standard incorporates a variable lower vapour pressure requirement
that ranges from 275 kPa(g) to 950 kPa(g) (European Committee for Standardisation, 2008).
This requirement is used to account for seasonal variations, and is set at a national level
96 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

(a) propane/propylene/n-butane. (b) propane/propylene/iso-butane.

(c) propane/n-butane/iso-butane. (d) propylene/n-butane/iso-butane.

Figure 4.12: Contour diagrams of RON for ternary mixtures of propane, propylene, n-
butane and iso-butane that comply with the MON and vapour pressure requirements
of the Australian Fuel Determination (Autogas) 2003 (Federal Register of Legislative
Instruments, 2003). Data is presented on a molar basis.
4.7 RONs of LPG satisfying existing fuel standards 97

(Table 4.8). Therefore, the minimum fuel vapour pressure may still vary considerably
from region-to-region during a given period. The Grade A specification is compulsory
during winter periods to ensure that sufficient fuel volatility exists for vehicles equipped
with vapour-phase fuel delivery systems. Outside of this period, many regions allow LPG
producers to comply with any of the five lower vapour pressure limits (e.g. the United
Kingdom). The upper vapour pressure limit is specified at 1550 kPa(g) and does not vary
with season. The minimum MON requirement is 89.0.

Table 4.8: EN 589 lower vapour pressure limits and corresponding grades (European
Committee for Standardisation, 2008). Limits are specified at a temperature of 40◦ C.

Grade Vapour pressure


kPa(g)
A 950
B 800
C 700
D 500
E 275

During non-winter periods, the wide range of allowable vapour pressures may lead to
significant variations in the composition of LPG in some regions. For example, the Grade
E specification would allow both pure n-butane and pure iso-butane (or mixtures thereof)
to be used as LPG. In the limiting case of n-butane, this fuel would have a RON of 93.5.
This value is significantly lower than pure propane (RON 109.4), which could be used as
LPG under any of the five seasonal grades. The Grade E specification would only prevent
the use of pure propylene as LPG, due to its considerably lower MON.

Since the Grade E specification only significantly restricts the propylene content in
LPG, the RONs for ternary mixtures of propane, propylene, n-butane and iso-butane
that satisfy the Grade D requirements are instead presented in Figure 4.13. Overall, this
fuel specification can be satisfied by a much broader range of fuel compositions than the
Australian LPG fuel standard, particularly those with significant butane content (Table
4.9).

The minimum RON of any fuel that complies with the vapour pressure and MON
98 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

(a) propane/propylene/n-butane. (b) propane/propylene/iso-butane.

(c) propane/n-butane/iso-butane. (d) propylene/n-butane/iso-butane.

Figure 4.13: Contour diagrams of RON for ternary mixtures of propane, propylene, n-
butane and iso-butane that comply with the MON and vapour pressure requirements of
the European EN 589 standard (Grade D) (European Committee for Standardisation, 2008).
Data is presented on a molar basis.
4.7 RONs of LPG satisfying existing fuel standards 99

Table 4.9: Fraction of the various ternary compositional spaces that satisfy the respec-
tive requirements of the Australian LPG fuel standard and the EN 589 Grade D fuel
specification.

Region of compliance (%)


Aust. EN 589
Compositional space (Autogas) Grade D
(a) propane, propylene, n-butane 28.4 52.9
(b) propane, propylene, iso-butane 56.3 82.0
(c) propane, n-butane, iso-butane 25.8 70.2
(d) propylene, n-butane, iso-butane 4.2 40.3

requirements of this fuel specification is 95.6 (Figure 4.13a). This value is approximately
4.0 octane numbers lower than that of any fuel that complies with the Australian LPG fuel
standard.

4.7.3 United States Propane HD-5

Propane HD-5 (Gas Processors Association, 1997), also known as Special-duty Propane
(ASTM International, 2011), is a US LPG fuel specification developed specifically for
use in spark-ignition engines. Fuels that comply with this specification contain at least
92.5 (% vol) propane. Compositional restrictions are also imposed on the maximum
concentrations of propylene and large hydrocarbons (C4 +) of 5 (% vol) and 2.5 (% vol)
respectively. Unsurprisingly, fuels that satisfy the Propane HD-5 specification are reported
to be far less variable in combustion characteristics than other LPG fuel specifications
(ASTM International, 2011).

The minimum RON for the Propane HD-5 specification is compared with those of the
alternative global LPG fuel specifications in Table 4.10. Propane is used as a reference
fuel, as this is the only pure constituent that would satisfy the requirements of each
standard. Predictably, Table 4.10 indicates that the minimum RON for Propane HD-5 is
comparable to pure propane. The minimum RON of 108.2 is significantly higher than the
corresponding value for the Australian LPG fuel standard (RON 99.5), along with the
most stringent European Grade A (winter) specification (RON 99.8).
100 The Research and Motor Octane Numbers of Liquefied Petroleum Gas

Table 4.10: Minimum RONs of fuels that satisfy the compositional or the MON/vapour
pressure requirements of various global LPG fuel standards.

EN 589
Pure HD-5 Aust. Grade Grade Grade Grade Grade
propane Propane (Autogas) A B C D E
RON 109.4 108.2 99.5 99.8 98.5 97.6 95.6 93.5

4.8 Implications for LPG fuel quality standards

This analysis has shown that LPG fuel quality standards impose significant restrictions on
fuel composition. Of particular note was that these standards currently constrain the RON
of LPG to be significantly higher than that of standard gasoline in all cases, whilst the
minimum MON is also significantly higher than that of most gasolines. This suggests that
relaxation of both the RON and MON requirements of LPG is possible. Indeed, Equations
4.6 and 4.7 showed that the lowest RON and MON of any LPG considered in this study
is above the RON and MON of standard gasolines. Therefore, either the RON or MON
limits for LPG may be able to be made redundant and replaced with a compositional
requirement, or that they could be set to those of standard gasolines.

Further, LPG fuelled engines which inject liquid fuel into the port or directly into the
cylinder are becoming more common (Mizushima et al., 2009; Watson and Phuong, 2007).
If this trend continues, such that liquid phase injection dominates the market, the lower
vapour pressure limit of LPG’s may also be reduced, or even eliminated. Of course, the
upper vapour pressure limit must remain as a mechanical requirement for fuel dispensing
and storage. However, further research is needed to establish the actual performance of
these fuels in modern production engines, before such recommendations can be made
with confidence. This could be part of a more substantial undertaking by the community,
which also includes re-activation and updating of LPG’s octane testing procedures.
4.9 Summary 101

4.9 Summary

This chapter presented an experimental study of the Research and Motor octane numbers
of LPG. A comprehensive set of RON and MON data for mixtures of propane, propylene, n-
butane and iso-butane was presented, using a method that is consistent with the currently
active ASTM Research and Motor test methods for liquid fuels. This data was then
compared to the limited data in the literature, and found to agree closely with equivalent
data in almost all cases, thus validating the test procedure used in this study.

Empirical models which relate LPG composition to its RON and MON were then
developed. This required statistical model reduction, such that the simplest relationships
between the constituent species’ mole fractions and the mixture octane number were
achieved. The model error was significantly less than an octane number in both cases, and
was also well within the reproducibility limits of the Research and Motor test methods
for liquid fuels. When the minimum MON and vapour pressure limits of typical fuel
standards were not imposed, the RON of LPG was found to exhibit non-linearity of up
to approximately 1.5 octane numbers, whilst the maximum non-linearity of the MON
was approximately 1.0 octane number. The non-linearity in the RON was primarily due
to propylene. Imposition of the European and Australian LPG fuel standards reduced
the degree of non-linearity in the RON marginally. Since the United States fuel standard
imposes a high propane content, such considerations were not relevant.

These observations were then argued to have implications for the future design of
LPG fuel quality standards. In particular, given the high RONs and MONs observed, the
RON or MON limits for LPG may be able to be made redundant and replaced with a
compositional requirement, or they could be set to those of standard gasolines. Further,
if the trend of liquid phase LPG injection becomes the dominant means of fuel delivery
in modern engines, the lower vapour pressure limit of LPG’s may also be reduced, or
even eliminated. However, the upper vapour pressure limit must remain as a mechanical
requirement for fuel dispensing and storage. Nonetheless, further research is needed to
establish the actual performance of these fuels in modern production engines, before such
recommendations can be made with confidence.
Chapter 5

Combustion Modelling and Analysis

5.1 Introduction

Chapter 2 stated that the final aim of this thesis was to establish ”the phenomena that need
to be included in an accurate model of LPG autoignition in a spark-ignition engine.” This is
a challenging aim since the CFR engine at standard knock conditions features several
simultaneous phenomena, including autoignition and knock, the piston motion, flame
propagation, quenching and incomplete combustion, heat losses and blow-by. All of
these phenomena couple to a greater or less extent. Thus, a comprehensive model is one
that includes all of them in a single simulation− something that is computationally very
challenging today.

This and the next chapter therefore take a different approach. They present an em-
pirically calibrated, two-zone combustion model of the CFR engine. This chapter exam-
ines sub-models of each of the non-kinetic phenomena listed above for each fuel using
in-cylinder pressure data measured at standard test conditions, but with autoignition
suppressed. The key parameters obtained from these models, namely the flame propa-
gation, the heat losses and the mass that is estimated to participate in autoignition, are
then discussed briefly. Finally, the simulated unburned zone temperature histories are
examined, suggesting that kinetic activity during compression and flame propagation may
be significant. This is confirmed in the following chapter, which takes these sub-models
and examines the detailed autoignition chemistry.

103
104 Combustion Modelling and Analysis

5.2 In-cylinder pressure data

The results presented in this chapter use in-cylinder pressure traces acquired in the
CFR engine. As was previously mentioned, this data was acquired with a Kistler 6125C
vibration compensated and heat shielded piezoelectric pressure transducer. This hardware
was installed within the cylinder head, in the location that normally accommodates the
ASTM Detonation pickup used to measure the octane number.

In-cylinder pressure data was acquired for propane, propylene, n-butane and iso-
butane, together with eight mixtures that span the likely combinations of these fuels
(Table 5.1). In most cases, the engine was operated in accordance with the standard
Research method test conditions. The dial indicator was therefore set to the barometrically
compensated value that corresponded to the measured RON of each fuel, as determined
in the proceeding chapter. The air-fuel ratio was also maintained at the value previously
identified to produce the maximum Knock Intensity on the ASTM Knockmeter (the so-
called Standard Knock Intensity (SKI) test condition). As a result, the measured in-cylinder
pressure traces correspond to the same operating conditions encountered during octane
rating. This data is used in the present chapter and subsequent chapter to analyse the
relationship between the autoignition behaviour during octane rating and the Research
octane numbers (RONs) of the different fuels.

5.2.1 Raw pressure traces with autoignition

Figure 5.1 presents the raw measured in-cylinder pressure traces and logarithmic pressure-
volume diagrams for the four pure constituents at their respective Research method SKI
test conditions. At these operating conditions, autoignition occurs for each fuel. This
condition can be recognised as the near instantaneous pressure increase that occurs in the
final stages of combustion (between the polytropic compression and expansion processes).
Consistent with the literature, this indicates the characteristic time of autoignition is very
small in comparison with that of the piston motion (Bradley and Morley, 1997; Konig and
Sheppard, 1990; Heywood, 1988).
5.2 In-cylinder pressure data 105

(a) Raw measured in-cylinder pressure versus crank angle.

(b) Logarithmic pressure-volume diagram.

Figure 5.1: Raw measured in-cylinder pressure traces and logarithmic pressure-volume
diagrams for propane, propylene, n-butane and iso-butane. Autoignition occurs for all
fuels, and is denoted for propane. Data is presented for the Research method SKI test
condition.
106 Combustion Modelling and Analysis

Table 5.1: Summary of the fuels examined in this chapter, along with the corresponding
Research method SKI test conditions.

Fuel composition (% mol) Research method SKI test condition

Dial
propane propylene n-butane iso-butane indicator (in.) RONa λaSKI

100.0 - - - 0.206 109.4 0.96


- 100.0 - - 0.356 100.2 1.03
- - 100.0 - 0.459 93.5 0.90
- - - 100.0 0.357 100.1 0.90
70.0 10.0 10.0 10.0 0.250 105.6 0.95
66.6 16.7 - 16.7 0.243 106.1 0.96
66.6 - 16.7 16.7 0.261 104.9 0.93
50.0 50.0 - - 0.291 103.2 0.98
50.0 - 50.0 - 0.351 100.4 0.91
41.7 16.6 41.7 - 0.345 100.7 0.92
16.7 - 66.6 16.7 0.418 96.6 0.90
- - 50.0 50.0 0.416 96.8 0.90
a measured values obtained during the work presented in Chapter 4.

The proceeding chapter demonstrated that the pure constituents have different RONs.
Therefore, the compression ratio corresponding to the SKI test condition must also be
varied in order to induce autoignition. Propane has the highest RON, and therefore also
has the highest compression ratio (RON, 109.4 and CR, 9.893:1). Intuitively, Figure 5.1
indicates that this causes the in-cylinder pressure to attain a higher value prior to the start
of combustion.

5.2.2 Criterion for selection of representative autoigniting pressure traces

Cycle-to-cycle variation occurs in all internal combustion engines, and results in varying
combustion and autoignition development (Figure 5.2). Given these variations, several
techniques have been proposed in the literature for selecting a single pressure trace that is
considered representative of a given operating condition. For non-autoigniting combustion,
some researchers have opted to match the average peak pressure and its location in the
crank angle domain, e.g. (Swarts et al., 2003; Lancaster et al., 1975). Other researchers
have found it useful to also include the net or gross indicated mean effective pressure
5.2 In-cylinder pressure data 107

(IMEP) in their selection criteria, e.g. (Abbasi Atibeh, 2012; Swarts, 2006). However, these
parameters are less suited to the analysis of autoignition, for which the autoignition timing
is clearly a key parameter.

Figure 5.2: Measured variation in the autoignition onset location for propane over 900
consecutive cycles. Data is presented for the Research method SKI test condition.

In the present work, a representative autoigniting pressure trace for each operating
condition was therefore selected on the basis of the autoignition behaviour. First, the
autoignition onset location, which is denoted in Figure 5.1a, was estimated for each of the
twelve fuels over 900 consecutive cycles using a weighted mean square error criterion
(Coetzer et al., 2006). The corresponding in-cylinder pressure at this location was also
computed. A single pressure trace that closely matched the mean values of these two
parameters over 900 cycles was then selected by inspection (Figure 5.3). There were no
additional constraints imposed on the combustion development prior to autoignition.
It can be argued that this approach is justified, given the correlation observed between
the combustion development and the autoignition location. This is consistent with data
reported elsewhere in the literature (Burluka et al., 2004).
Combustion Modelling and Analysis
Figure 5.3: Measured raw in-cylinder pressure traces for propane, propylene, n-butane and iso-butane. Data is presented for the
cycles corresponding to the least and most advanced combustion development, together with the representative cycle for each

108
operating condition. All data was acquired at the Research method SKI test condition, from a batch of 900 consecutive cycles.
5.2 In-cylinder pressure data 109

5.2.3 Raw pressure traces without autoignition

It is difficult to determine the crank angle resolved mass fraction burned (MFB) history
from in-cylinder pressure traces that exhibit autoignition. Rather, this parameter can
more easily be estimated using in-cylinder pressure traces where the charge is fully
consumed by turbulent flame propagation. To determine the pressure development that
would occur at the SKI test condition in the absence of autoignition, a small amount of
dilute tetraethyllead (TEL) was added to the intake charge. The experimental apparatus
and technique used to prepare these leaded fuel mixtures is described in Section 3.6.1.
Consistent with data reported in the literature (Bradley and Morley, 1997; Ellison et al.,
1968; Pahnke et al., 1954; Downs et al., 1951), the addition of a small amount of TEL fully
suppressed the autoignition of each of the twelve fuels examined.

The use of TEL to suppress autoignition was considered preferable to several al-
ternative techniques that were also considered. For example, the in-cylinder pressure
development in the absence of autoignition could also be estimated by reducing the com-
pression ratio, retarding the spark timing and/or varying the air-fuel ratio. However,
these parameters alter the mixture temperature, pressure and residual gas fraction, and
therefore the laminar flame speed (Lindstrom et al., 2005). Rather, if the addition of the
TEL led to similar pressure development histories prior to autoignition, then it could
reasonably be concluded that it only suppressed autoignition, and does not affect flame
propagation. This is consistent with the general consensus in the literature as to how TEL
can suppress autoignition and not affect flame propagation significantly (Bradley and
Morley, 1997; Pitz and Westbrook, 1986; Kirsch and Pye, 1975; Ellison et al., 1968; Salooja,
1965; Pahnke et al., 1954; Downs et al., 1951).

The representative autoigniting pressure traces selected for each of the pure con-
stituents, together with the corresponding non-autoigniting pressure traces, are presented
in Figures 5.4 and 5.5. The latter was selected by evaluating the residual sum of squares
(RSS) between the representative autoigniting pressure trace prior to autoignition, and
each of the 900 non-autoigniting pressure traces acquired when the engine was operated
on the gaseous fuel that contained TEL. The non-autoigniting pressure trace that provided
the minimum RSS of the pressure between the spark and autoignition locations was
Combustion Modelling and Analysis
Figure 5.4: Representative autoigniting and non-autoigniting pressure traces for propane, propylene, n-butane and iso-butane.
110
Data is presented for the Research method SKI test condition.
5.2 In-cylinder pressure data

Figure 5.5: Representative autoigniting and non-autoigniting logarithmic pressure-volume diagrams for propane, propylene,
n-butane and iso-butane. Autoignition is denoted for n-butane. Data is presented for the Research method SKI test condition.
111
112 Combustion Modelling and Analysis

selected. Figures 5.4 and 5.5 show that the agreement between the two pressure traces
is good for each of the pure constituents, with the only significant departure being the
autoignition event itself. Similar agreement was also obtained for the eight mixtures of
these constituents examined in this chapter. Further, Figure 5.6 shows that the addition of
TEL to the gaseous fuel also does not appear to affect the cycle-to-cycle variations. This
data was acquired at sub-critical compression ratio conditions, where neither the leaded or
gaseous fuel exhibited autoignition. This trend is consistent with data reported elsewhere
in the literature (Ellison et al., 1968), and provides further evidence that TEL only suppress
autoignition, whilst not significantly affecting flame propagation.

5.3 Engine-out emissions with autoignition suppressed

The measured engine-out emissions data was used to determine the combustion efficiency
for each fuel. This information is used to estimate the fraction of the starting fuel burned
within the cylinder. The latter is provided as an input to the combustion model developed
in the subsequent sections of this chapter. As mentioned previously, the engine-out exhaust
emissions were measured with an Auto-diagnostics ADS 9000P five gas analyser. This
hardware measured the dry exhaust concentrations of oxygen (O2 ), carbon monoxide (CO),
carbon dioxide (CO2 ), nitric oxide (NO) and the total unburned hydrocarbons (UHCs).
The wet species concentrations were computed from the dry measurements using the
method described by Heywood (Heywood, 1988). This data is summarised in Table 5.2
and Figure 5.7.

Consistent with the literature, Figure 5.7 indicates that the CO emissions correlate
strongly with lambda, and are largely independent of fuel composition (Heywood, 1988).
NO also has a strong dependence on lambda, but instead increases with increasing
lambda. For most fuels, the measured NO emissions typically range between 800-1600
ppm. By contrast, pure propylene has considerably higher NO emissions of approximately
2500 ppm, owing to its lean mixture (λ=1.03) in the Research method. These trends are
directionally consistent with those reported elsewhere in the literature, which typically
show a considerable increase in NO emissions between λ=1.00 and λ=1.10 (Hilliard and
5.3 Engine-out emissions with autoignition suppressed 113

Figure 5.6: Representative leaded and unleaded pressure traces for propane and n-butane
at sub-critical compression ratio test conditions, together with the upper and lower vari-
ability limits observed in each batch of 900 cycles. The amount of TEL added to the intake
is reported as a percentage of the total fuel mass entering the engine.
114 Combustion Modelling and Analysis

Table 5.2: Measured engine-out exhaust emissions of CO, total UHCs and NO (wet basis)
for the twelve fuels with autoignition suppressed. Data is presented for the Research
method SKI test condition.

Fuel composition (% mol.) Engine-out exhaust emissions


total NO
CO
UHCs
propane propylene n-butane iso-butane λSKI CR (% mol.) (ppm C1 ) (ppm)

100.0 - - - 0.96 9.893:1 1.70 912 1530


- 100.0 - - 1.03 7.860:1 0.20 222 2410
- - 100.0 - 0.90 6.929:1 3.69 720 792
- - - 100.0 0.90 7.849:1 3.65 700 890
70.0 10.0 10.0 10.0 0.95 9.182:1 2.00 765 1355
66.6 16.7 - 16.7 0.96 9.287:1 1.70 792 1530
66.6 - 16.7 16.7 0.93 9.021:1 2.64 822 1190
50.0 50.0 - - 0.98 8.700:1 1.16 540 1610
50.0 - 50.0 - 0.91 7.912:1 3.33 750 920
41.7 16.6 41.7 - 0.92 7.977:1 3.20 685 990
16.7 - 66.6 16.7 0.90 7.267:1 3.69 765 855
- - 50.0 50.0 0.90 7.285:1 3.62 722 875

Wheeler, 1979; Sakai et al., 1973; Quader, 1973; Benson and Stebar, 1971; Nebel and Jackson,
1958; Hanson and Egerton, 1937). A correlation for the total UHCs across all fuels is less
clear.

Combustion efficiency

The combustion efficiency, ηc , can be defined as,

∑i=1 xi Q HV, i
ηc = 1 −   ,
ṁ f / ṁ a + ṁ f Q HV, f (5.1)

where xi and Q HV, i denote the mass fraction and lower (net) heating value of species i
in the exhaust respectively, ṁ a and ṁ f denote the mass flow rate of air and fuel through
the engine respectively, and Q HV, f denotes the lower heating value of the fuel. The
5.3 Engine-out emissions with autoignition suppressed 115

Figure 5.7: Measured engine-out exhaust emissions for CO, NO and total UHCs (wet
basis) versus lambda and compression ratio. Data is presented for the Research method
SKI test condition.
116 Combustion Modelling and Analysis

combustion efficiency in terms of the measured CO and the total UHC emissions can
therefore be expressed as,

( xCO Q HV, CO + xUHCs Q HV, UHCs )


ηc = 1 −   .
ṁ f / ṁ a + ṁ f Q HV, f (5.2)

The lower heating value of CO is 10.1 MJ/kg (Heywood, 1988). However, the lower
heating value of the UHCs must be estimated, since only the total concentration is known
and not its composition. In this analysis, the UHCs were assumed to be completely
unreacted, and therefore the lower heating value corresponded to that of the fuel (Table
5.3).

Table 5.3: Lower heating values for the pure constituents (Turns, 2000), as used to calculate
the combustion efficiency.

Constituent Q HV (MJ/kg)
propane 46.4
propylene 45.8
n-butane 45.7
iso-butane 45.6

The combustion efficiency for the various fuels is presented in Figure 5.8. Unsurpris-
ingly, the combustion efficiency shows a strong dependence on lambda, and decreases
with decreasing lambda. Propylene has the highest combustion efficiency, owing to its
marginally lean condition. Conversely, the combustion efficiency for the butane-rich fuels
generally does not exceed 90%. This dependence of the combustion efficiency on lambda
is consistent with data reported elsewhere in the literature (Heywood, 1988; Lambert, 1984;
Stivender, 1971; D’Alleva and Lovell, 1936).

These results also indicate that CO is the largest contributor to the combustion inef-
ficiency. In general, the inclusion of total UHC emissions in Equation 5.2 reduces the
combustion efficiency by less than half a percent from that determined based on only the
measured CO emissions. It can therefore be argued that the uncertainty in the measured
5.4 Two-zone engine model without autoignition 117

Figure 5.8: Combustion efficiency for the twelve fuels with autoignition suppressed versus
lambda. Data is presented for the Research method SKI test condition.

total UHC emissions, along with the need to estimate the lower heating value, does not
significantly affect the combustion efficiency.

5.4 Two-zone engine model without autoignition

The simulation tool GT-Power (Gamma Technologies Inc., 2012a) is now used to develop
a model of the CFR engine that can be used to analyse the non-autoigniting combustion
behaviour of the various fuels. The full engine model is comprised of sub-models of the
intake and exhaust systems, along with the engine itself. The model is first validated using
empirical data, and is then used to analyse the combustion behaviour of the twelve fuels
at octane rating conditions.

In this chapter, the software is exclusively operated in GT-Power’s so-called reverse-run


combustion simulation mode. The reverse-run combustion simulation calculations are
performed for the closed portion of the cycle (i.e. IVC to EVO) using the two-zone model
of the combustion process presented in Figure 5.9. Several assumptions are adopted in
118 Combustion Modelling and Analysis

the model in order to simplify the governing equations. First, the pressure within the
combustion chamber is assumed to be uniform at a given crank angle. This is based on the
assumption that the characteristic time of sound wave propagation across the combustion
chamber is far smaller than the characteristic time of combustion. This assumption is
generally valid until shortly after autoignition (Jenkin et al., 1997; Moses et al., 1995; Witze
and Green, 1993; Konig and Sheppard, 1990). Additionally, heat exchange between the
unburned and burned zones is not modelled. However, each zone may exchange heat
with the combustion chamber surfaces.

Figure 5.9: The two-zone engine model used in the reverse-run combustion simulation.

Prior to combustion, the contents of the cylinder are all assumed to exist within the
unburned zone. Once combustion has commenced, mass is transported from the unburned
zone to the burned zone. The amount of mass transported at each time step dictates the
rate of flame propagation, and therefore the unburned zone pressure-temperature history.
These calculations are performed iteratively based on the conservation of mass and energy,
with the intention of matching the simulated in-cylinder pressure history with that of the
representative non-autoigniting pressure trace. The final fraction of the fuel mass burned
within the cylinder is estimated based on the combustion efficiency (refer to Section 5.3),
which is also provided as an input to the model.
5.4 Two-zone engine model without autoignition 119

Chemical equilibrium calculations are performed for the burned zone at each time step.
This yields the individual species concentrations of the products of combustion, which by
default include CO2 , H2 O, N2 , O2 , CO, H2 , H, O, OH, NO and N (Gamma Technologies
Inc., 2012a). The thermodynamic properties used in these calculations are taken from the
NASA Glenn thermodynamic database (NASA Glenn Research Center, 2006). All gases
are assumed ideal, with the the specific heat, enthalpy and entropy each evaluated using
a nine term polynomial fit of the data. The individual species concentrations within the
burned zone are then used to compute the internal energy by summation over all species.
The updated burned and unburned zone temperatures can then be computed by applying
conservation of mass and energy.

5.4.1 Model formulation and calibration

The full engine model implemented in GT-Power is presented in Figure 5.10. The model
was comprised of the combustion chamber and crank-train objects, together with sub-
models of the intake and exhaust systems. The intake and exhaust sub-models are first
validated using empirical data, and are then used to determine the state of the fresh
air-fuel mixture at intake valve closure. The geometry of all components was based on
that of the standard CFR engine, and included the plain cylindrical combustion chamber
and side-mounted spark-plug.

Table 5.4: CFR engine geometry applied in the GT-Power model.

Attribute Value

Bore 82.5 mm
Stroke 114.3 mm
Conrod length 254.0 mm
Head/bore area ratio 1
Piston/bore area ratio 1

Each of the sub-models implemented within the full engine model is described in the
following section of the chapter. The calibration and validation of the full engine model
Combustion Modelling and Analysis
Figure 5.10: GT-Power model of the CFR engine (left to right: intake system, carburettor venturi, engine cylinder and crankcase,
120
exhaust system).
5.4 Two-zone engine model without autoignition 121

using experimental data is described thereafter.

Gas exchange sub-models

In order to simulate the state of the air-fuel mixture at intake valve closure, the full engine
model was equipped with sub-models of the intake and exhaust gas exchange processes.
These sub-models contained empirical forward and reverse discharge coefficient data
for the standard intake and exhaust valves. This data was obtained from the literature
(Vancoillie, 2013). The valve lift profiles were measured in-house using a dial indicator
(Appendix B).

The gas exchange sub-models are assumed to be the only mechanism by which mass is
transported across the cylinder boundaries. Thus, leakage of the air-fuel charge between
the cylinder liner and piston rings (so-called blow-by) was neglected. Further, the so-called
engine trapping efficiency was assumed to be 100%, i.e. none of the fresh air-fuel mixture
inducted into the engine blows through into the exhaust during the valve overlap period.
The trapping efficiency cannot readily be measured on a standard engine, but is typically
very close to 100% for most 4-stroke engines (Heywood, 1988). Moreover, this assumption
is considered particularly relevant to the CFR engine, which has a very short valve overlap
duration of 5.0 crank angle degrees (ASTM International, 2012a,b). The validity of these
assumptions is subsequently examined in Section 5.4.2.

Liquid fuel carburettor sub-model

The liquid fuel carburettor was not utilised when the engine was operated on the gaseous
fuels. However, due to the significant flow restriction imposed by the venturi section
(Leppard, 1990; Johnson et al., 1965), it is still necessary to accurately model this hardware
in order to correctly simulate the volumetric efficiency of the engine.

The liquid fuel carburettor was modelled as a simple venturi, with the key geometric
features matching those of the standard hardware fitted to the CFR engine (Figure 5.11).
The converging section was modelled as a 28 mm diameter pipe with an adjoining bell-
122 Combustion Modelling and Analysis

Figure 5.11: Modelling approach used to represent the liquid fuel carburettor in GT-Power,
together with a sectional view of the actual hardware on the CFR engine.

mouth to atmosphere. The forward and reverse discharge coefficients for the bellmouth
were assumed to be unity. The diverging section of the carburettor was modelled as a
70 mm tapered pipe, with an inlet and exit diameter of 13.6 mm and 30 mm respectively.
The rear-most section of the carburettor was modelled as a round pipe with an internal
diameter of 30 mm.

Fuel delivery and air-fuel ratio specification sub-model

Consistent with the operation of the modified CFR engine, the gaseous fuel was introduced
into the intake air upstream of the liquid fuel carburettor (Figure 5.10). The temperature
and vapour fraction of all gaseous fuels was assumed to be 298 K and unity respectively.

The fuel flow rate for each operating condition is evaluated by the code based on
5.4 Two-zone engine model without autoignition 123

the simulated volumetric efficiency, together with a user-imposed value of lambda. The
volumetric efficiency is first determined by running the full engine model until a steady-
state solution is attained. Convergence was assumed to have been achieved once the
volumetric efficiency did not vary by more than 0.2% across three consecutive cycles.

External engine component temperatures

The intake and exhaust system components, together with the ports, were each assigned a
fixed wall temperature in order to simulate the heat transfer that occurred between these
components and the working fluid. The convective heat transfer coefficient is evaluated
by the code for each component using convective heat transfer correlations for turbulent
flow (Gamma Technologies Inc., 2012b).

The air intake pipe is thermally insulated between the standard air intake heater and
the liquid fuel carburettor. Therefore, the temperature of the components in this section
of the model were assumed to be in thermal equilibrium with the intake air. The latter
temperature is prescribed at 325 K in the Research method. The liquid fuel carburettor is
mounted between the intake pipe and the cylinder head, and is therefore assumed to be at
a marginally higher temperature of 340 K. The cylinder head is assumed to be in thermal
equilibrium with the cooling water jacket, which is maintained at a temperature of 373 K
in the Research method.

The intake and exhaust ports were assumed to be at a temperature of 373 K and 500 K
respectively (Caton and Heywood, 1981; Danis, 1973). The latter temperature was also
used for the section of the exhaust immediately downstream of the port, but upstream of
the water-cooled section of the exhaust. All other components in the model were assumed
to be at a temperature of 300 K.

In-cylinder heat transfer sub-model

Forced convection is the primary mechanism through which heat is transferred from
the working fluid to the combustion chamber surfaces in an internal combustion engine
124 Combustion Modelling and Analysis

(Heywood, 1988). However, the overall convective heat transfer is unable to be measured
in a standard engine, and therefore must be estimated. This is usually undertaken on
an instantaneous, spatially-averaged basis using Newton’s law of cooling. The heat flux
across the boundary between the working fluid and the combustion chamber surfaces can
be expressed as,

dQ 
= hc A Tgas − Twall ,
dt (5.3)

where hc denotes the convective heat transfer coefficient and A denotes the contact area
between the working fluid and combustion chamber surfaces. The terms Tgas and Twall
represent the temperatures of the working fluid and combustion chamber surfaces respec-
tively. The convective heat transfer coefficient is frequently evaluated using the Woschni
correlation without swirl (Woschni, 1967),

hc = ζ B −0.2 p 0.8 T −0.55 w 0.8 ,


(5.4)

where ζ denotes a convective heat transfer coefficient multiplier, B denotes the cylinder
bore, and w denotes the mean gas velocity. In a 4-stroke engine without swirl, the mean
gas velocity is defined as (Woschni, 1967),

 
Vs T1
w = C1 S¯p + C2 ( p − p0 ) ,
p1 V1 (5.5)

where C1 and C2 are constants, S¯p denotes the mean piston speed, and Vs denotes the
swept volume. The terms p1 , T1 and V1 correspond to the pressure, temperature and
volume of the working fluid at a reference state, respectively. Finally, p represents the
5.4 Two-zone engine model without autoignition 125

instantaneous pressure of the working fluid, and p0 the corresponding pressure of the
working fluid under motoring conditions.

Considerable evidence in the literature indicates that calibration of the Woschni heat
transfer model is necessary in order to reproduce measured in-cylinder pressure data,
e.g. (Baratta and Spessa, 2011; Bos, 2007; Baratta et al., 2005; Catania et al., 2003, 2000;
Guezennec and Hamama, 1999). This calibration can be undertaken by adjusting the value
of the Woschni convective heat transfer coefficient multiplier, ζ, which has a default value
of unity (Equation 5.4).

To calibrate the convective heat transfer coefficient multiplier, Equation 5.3 indicates
that it is first necessary to assign a temperature to the chamber surfaces. These include the
piston crown, the cylinder head and the cylinder wall temperatures. These temperatures
cannot readily be measured in a standard engine, and therefore must be estimated. In
the present work, these temperatures were estimated using experimental data reported
in the literature for the CFR engine. The cylinder head and piston crown temperatures
were assumed to be 440 K (Imming, 1944; Wittmann and Smith, 1946; Marr et al., 2010). In
general, the temperature of the cylinder wall is considered to be marginally lower than
that of the piston crown (Rahman et al., 2010; Heywood, 1988), enabling heat flux from
the piston to the bore. Therefore, the cylinder wall temperature was assumed to be 410
K (Demuynck et al., 2009; Nishiwaki, 1998; Lyford-Pike and Heywood, 1984; Deslandes,
1975).

Parametric variation of the simulated pressure traces estimated from these tempera-
tures and the ζ values can then be compared to the experimental pressure traces. Figure
5.12 shows the sum of squared errors (SSE) between these modelled and measured pres-
sure traces with the cylinder head/piston temperature and the cylinder wall temperature
held constant at 440 K and 410 K respectively. This data shows the associated optimal
values of the convective heat transfer coefficient multiplier for these temperatures are
approximately constant across all test cases (1.10-1.15). In general, Table 5.5 indicates the
same behaviour was also observed for the eight mixtures of these fuels. Collectively, this
behaviour is consistent with several other studies in the literature, which indicate the
Woschni correlation without swirl marginally under-predicts the heat transfer that occurs
126 Combustion Modelling and Analysis

in real engines (Irimescu, 2013; Hamada et al., 2010; Soyhan et al., 2009; Chang et al., 2004;
Borman and Nishiwaki, 1987).

Table 5.5: Optimised Woschni convective heat transfer coefficient multiplier (ζ) for the
twelve fuels with autoignition suppressed.

Fuel composition (% mol)


propane propylene n-butane iso-butane λSKI CR ζ

100.0 - - - 0.96 9.893:1 1.14


- 100.0 - - 1.03 7.860:1 1.15
- - 100.0 - 0.90 6.929:1 1.14
- - - 100.0 0.90 7.849:1 1.10
70.0 10.0 10.0 10.0 0.95 9.182:1 1.07
66.6 16.7 - 16.7 0.96 9.287:1 1.08
66.6 - 16.7 16.7 0.93 9.021:1 1.11
50.0 50.0 - - 0.98 8.700:1 1.21
50.0 - 50.0 - 0.91 7.912:1 1.07
41.7 16.6 41.7 - 0.92 7.977:1 1.08
16.7 - 66.6 16.7 0.90 7.267:1 1.12
- - 50.0 50.0 0.90 7.285:1 1.10

Estimated start and end of combustion

As previously shown in Figure 5.1, the compression and expansion processes in an internal
combustion engine are well approximated as a polytropic process (Heywood, 1988),

p V n = constant,
(5.6)

where p denotes the pressure, V denotes the cylinder volume, and n denotes the polytropic
index. Since combustion represents a departure from polytropic behaviour, the pressure
trace itself can be used to estimate the start and end of combustion. Figure 5.13 presents
the derivative of the logarithm of the pressure with respect to the logarithm of the volume.
The estimated start and end of combustion are readily identified, since rearrangement of
5.4 Two-zone engine model without autoignition 127

(a) propane. (b) propylene.

(c) n-butane. (d) iso-butane.

Figure 5.12: Absolute error (% IMEP) between the simulated pressure history and the rep-
resentative non-autoigniting pressure trace versus the simulated convective heat transfer
coefficient multiplier. Data is presented for the Research method SKI test condition, with
the cylinder head/piston and cylinder wall temperatures held constant at 440 K and 410 K
respectively.
128 Combustion Modelling and Analysis

Equation 5.6 yields,

d (log p)
n = − .
d (log V ) (5.7)

The horizontal lines in Figure 5.13 correspond to the polytropic index for the compression
and expansion strokes respectively. In all cases, this value is approximately 1.4, which
is approximately equal to the ratio of specific heats for air, as expected. Figure 5.13 also
indicates that the estimated start of combustion is similar for each fuel, owing to the
constant spark timing used in the Research method for all fuels. The estimated end of
combustion generally varies by a few crank angle degrees.

Flame propagation sub-model

The representative non-autoigniting pressure traces were used to generate a crank angle
resolved burn rate profile for each fuel. As previously discussed, this involves operating
the code in the so-called reverse-run combustion simulation mode. This analysis considers
only the portion of the cycle where both valves are closed, and therefore the initial
conditions inside the cylinder must be estimated using the full engine model (Figure 5.10).
The parameters that must be estimated include the volumetric efficiency, the residual
fraction, the trapped fuel mass, along with the initial fuel vapour fraction.

Initially, the measured engine-out exhaust emissions are used to determine the com-
bustion efficiency (refer to Section 5.3). This information is used to determine the fraction
of the starting fuel that is burned within the cylinder. The crank angle by which this fuel
must be burned is imposed based on the estimated end of combustion. Subject to these
constraints, the two-zone model of the combustion process is then used to determine the
appropriate amount of mass that must be transferred between the unburned and burned
zones at each time step, in order to reproduce the representative non-autoigniting pressure
trace. This procedure is dependent upon the convective heat transfer between the working
fluid and combustion chamber surfaces, and is therefore iterative.
5.4 Two-zone engine model without autoignition 129

(a) propane. (b) propylene.

(c) n-butane. (d) iso-butane.

Figure 5.13: Estimated start (•) and end (◦) of combustion for propane, propylene, n-
butane and iso-butane. Data is presented for the Research method SKI test condition.
130 Combustion Modelling and Analysis

5.4.2 Model validation

The intake and exhaust sub-models were validated using the steady-state fuel flow data
and lambda measurements acquired during the engine experiments. As was previously
discussed, these sub-models are used in conjunction with the gas exchange sub-models to
simulate the volumetric efficiency of the engine. The steady-state fuel flow rate delivered
to the engine is then evaluated by the code, making use of the user-imposed relative
exhaust air-fuel ratio (lambda). The simulated fuel flow rate therefore depends on the
simulated volumetric efficiency, and can be compared with the experimental fuel flow
rate data as a test of the gas exchange modelling.

The simulated volumetric efficiency for each operating condition is presented in Figure
5.14. The mean simulated value is 79.8%, and does not show any dependence on either
lambda or compression ratio. Normally, the volumetric efficiency for pre-vapourised
fuels would be expected to increase with compression ratio (Heywood, 1988; Sparrow,
1925). However, several studies have shown that the effect of the compression ratio on the
volumetric efficiency is small for CFR engines (Tribbett et al., 2010; Gallas, 1959). Consistent
with the simulation data, these studies also indicate that small lambda variations have
only a minor effect on the volumetric efficiency.

The mean simulated volumetric efficiency obtained in this study is approximately


3-5% lower than that measured with a CFR engine operated under similar conditions on
liquid hydrocarbon fuels (Tribbett et al., 2010). The higher volumetric efficiency obtained
in the latter study can likely be attributed to the thermodynamic charge cooling effect
that arises from the evaporation of the fuel. This effect considerably reduces the intake
mixture temperature in the Research method for all liquid fuels (Foong et al., 2013).
Unsurprisingly, the experimental data presented in Figure 5.15 indicates that intake charge
cooling is not present for the pre-vapourised fuels examined in the present work. Given
these considerations, the marginally lower simulated volumetric efficiencies obtained in
this study relative to those observed experimentally for liquid hydrocarbons fuels appears
reasonable.

The simulated steady-state fuel flow rates are compared with the measured steady-
state fuel flow rates in Figure 5.16. In general, the two values show good agreement.
5.4 Two-zone engine model without autoignition 131

Figure 5.14: Volumetric efficiency versus compression ratio. Data is presented for the
Research method SKI test condition.

Figure 5.15: Measured engine intake temperature under motoring (air only) and firing (air
and fuel) conditions (refer to Figure 3.4b). Data is presented for the Research method SKI
test condition.
132 Combustion Modelling and Analysis

Figure 5.16: Measured steady-state fuel flow rate versus simulated steady-fuel flow rate.
Data is presented for the Research method SKI test condition.

However, in all cases the simulated steady-state fuel flow rate is approximately 2-3%
higher than the measurements, which is a systematic bias opposite to that caused by
charge cooling. This could indicate the simulated volumetric efficiency is marginally
higher than the actual value, or alternatively that mass is removed from the combustion
chamber in the engine experiments, after intake valve closure.

In practice, leakage of the air-fuel charge between the cylinder liner and piston rings (so-
called blow-by) can contribute to the latter, and this has been ignored in these simulations.
Measurements undertaken in CFR engines indicate that blow-by is typically between
1-2% of the total charge per cycle (Rauckis and McLean, 1979; Deslandes, 1975), which
is directionally consistent with the observed behaviour in Figure 5.16. Thus, blow-by is
likely the main cause of this discrepancy, but since the errors are small, was not included
in the model.

The simulated cycle-averaged mixture intake temperature under firing conditions is


compared with the measured value obtained from the engine experiments in Figure 5.17.
Consistent with the measurement of this parameter on the engine, the simulated value
was obtained from a modelled thermocouple positioned immediately upstream of the
5.4 Two-zone engine model without autoignition 133

Figure 5.17: Measured engine intake temperature under firing conditions versus simulated
cycle-averaged intake temperature under firing conditions. Data is presented for the
Research method SKI test condition.

intake port (refer to Figure 5.10). In general, the two values differ by less than 2 K. This
suggests that the temperatures assigned to the intake system components, along with
the heat transfer sub-models within these components, are reasonable. Collectively, the
results presented in this section suggest the full engine model is accurately capturing the
fuel and air flow through the CFR engine.

The resulting simulated pressure histories obtained using the calibrated heat transfer
sub-models are compared with the representative non-autoigniting pressure traces in
Figure 5.18. For each fuel and operating condition, the agreement between the two
pressure histories is good. This suggests the overall modelling approach is reasonable.
134 Combustion Modelling and Analysis

Figure 5.18: Measured and reverse-run combustion simulation pressure histories versus
crank angle. Data is presented for the Research method SKI test condition.
5.4 Two-zone engine model without autoignition 135

Figure 5.18: Measured and reverse-run combustion simulation pressure histories versus
crank angle. Data is presented for the Research method SKI test condition.
136 Combustion Modelling and Analysis

Figure 5.18: Measured and reverse-run combustion simulation pressure histories versus
crank angle. Data is presented for the Research method test condition with lambda not
corresponding to the value for SKI.
5.5 Analysis of modelling results 137

5.5 Analysis of modelling results

5.5.1 Mass fraction burned histories during octane rating

The resulting, modelled mass fraction burned (MFB) histories for the various fuels with
autoignition suppressed are presented in Figure 5.19. The 50% MFB location and 10-90%
MFB duration of the various fuels is presented in Figure 5.20. For each profile, the crank
angle at which autoignition was observed during octane rating has been superimposed
on the data. Propane has the fastest burn rate at the SKI test condition, whilst iso-butane
has the slowest. The burn rate for propylene at the SKI test condition is also slower than
that of propane. This behaviour is at odds with the laminar flame speed measurements
for these constituents (Bechtold and Matalon, 2001; Davis and Law, 1998), but can be
attributed to the higher compression ratio used for propane in the standard octane tests
(due to its higher RON). This higher compression ratio also leads to a smaller residual
gas fraction, and therefore less mixture dilution relative to propylene. Higher mixture
dilution is known to cause a reduction in the laminar flame speed (Razus et al., 2010, 2009;
Lindstrom et al., 2005; Heywood, 1988).

The three mixtures that contain significant propane content (≥66.0%) exhibit near
identical MFB histories and autoignition locations (Figure 5.19b). These fuels also have
very similar RONs. Similar behaviour is also observed for the two mixtures in Figure
5.19c that contain ≥42.0% n-butane (RON 100.4 and 100.7), together with both mixtures in
Figure 5.19d (RON 96.6 and 96.8). This suggests that the flame propagation is relatively
insensitive to variations in the minor constituents, which is consistent with the laminar
flame speed variations for propane/n-butane/iso-butane mixtures (Mizushima et al., 2010;
Razus et al., 2010; Liao et al., 2005, 2004).

5.5.2 Mass fraction burned at autoignition

The estimated mass fraction burned at the onset of autoignition is presented in Figure 5.21.
This property was inferred from the non-autoigniting data presented in Figure 5.19, and
the crank angle at which autoignition was observed during octane rating. Unsurprisingly,
138 Combustion Modelling and Analysis

Figure 5.19: Mass fraction burned histories obtained from the reverse-run combustion
simulation without autoignition, together with the location of the autoignition onset
during octane rating. Data is presented for propane (Pr), propylene (Pryl), n-butane (nB),
iso-butane (iB) and mixtures thereof at the Research method SKI test condition.
5.5 Analysis of modelling results 139

Figure 5.20: 50% mass fraction burned crank angle and 10-90% burn duration. Data is
presented for propane (Pr), propylene (Pryl), n-butane (nB), iso-butane (iB) and mixtures
thereof at the Research method SKI test condition.

Figure 5.22 also indicates the MFB at autoignition correlates somewhat with the RON.
In this case, the fuels with higher RONs tend to have a smaller unburned mass fraction
participating in autoignition, and vice-versa. These results also compare reasonably with
the data reported by Yates and co-workers for PRFs with similar octane numbers (Yates
and Viljoen, 2008; Swarts et al., 2004).

5.5.3 Unburned zone temperature histories

The simulated unburned zone temperature histories for the pure constituents and the
various mixtures with autoignition suppressed are presented in Figures 5.23 and 5.24
respectively. These histories are presented up to the crank angle at which autoignition was
observed during octane rating. Figure 5.23 indicates that propane has the highest peak
unburned zone temperature (approximately 930 K). This can be attributed to its higher
compression ratio in the standard octane tests, which also results in propane exhibiting a
140 Combustion Modelling and Analysis

Figure 5.21: Estimated mass fraction burned at the onset of autoignition obtained from
the reverse-run combustion simulation with chemical equilibrium. Data is presented for
propane (Pr), propylene (Pryl), n-butane (nB), iso-butane (iB) and mixtures thereof at the
Research method SKI test condition.

Figure 5.22: Mass fraction burned at autoignition versus RON. Data is presented for the
Research method SKI test condition.
5.6 Summary 141

Figure 5.23: Simulated unburned zone temperature versus crank angle for the pure
constituents. Data is presented for the Research method SKI test condition.

significantly higher mixture temperature, even prior to the start of combustion. The other
pure constituents have peak unburned zone temperatures of approximately 900 K.

For the majority of the fuels examined, the crank angle at which the simulated peak
unburned zone temperature occurs does not correspond to that at which autoignition was
observed during octane rating (Figure 5.25). This is likely the result of two factors. First,
this equilibrium model of course does not permit any kinetic activity in the unburned
zone during the flame propagation. Further, post-top dead centre, the expansion stroke
results in the unburned gas temperature falling as the charge is consumed. The inclusion
of kinetic modelling is now the focus of the next chapter.

5.6 Summary

This chapter presented an experimental and modelling study of the combustion behaviour
of twelve LPG fuels during octane rating. In-cylinder pressure data was acquired for each
fuel in the CFR engine with a piezoelectric pressure transducer. This data was acquired
for autoigniting and non-autoigniting operating conditions. The latter was acquired
by adding a small amount of dilute tetraethyllead to the intake, which suppressed the
142 Combustion Modelling and Analysis

Figure 5.24: Simulated unburned zone temperature versus crank angle for the eight
mixtures. Data is presented for mixtures containing propane (Pr), propylene (Pryl), n-
butane (nB) and iso-butane (iB) at the Research method SKI test condition.
5.6 Summary 143

Figure 5.25: Experimental autoignition location during octane rating versus simulated
peak unburned zone temperature location. Data is presented for the Research method SKI
test condition.

autoignition that normally occurred during octane rating, without significantly affecting
the flame propagation. As such, the non-autoigniting results presented in this chapter,
namely the flame propagation and in-cylinder heat transfer, along with the sub-models
that describe the state of the in-cylinder mixture at intake valve closure, can be used in the
kinetic modelling presented in the next chapter.

Representative non-autoigniting pressure traces were selected for each operating


condition, and were used as inputs to a combustion model. The full model was comprised
of the standard CFR engine, together with sub-models of the intake and exhaust systems.
The full model was able to simulate the air and fuel flow through the engine, and showed
good agreement with steady-state measurements. The closed portion of the cycle assumed
chemical equilibrium, and was calibrated for each fuel using the representative non-
autoigniting pressure traces and measured engine-out emissions data. The latter was used
to determine the fraction of the total fuel that was burned within the cylinder.

Calibration of the in-cylinder convective heat transfer sub-model was then undertaken.
Consistent with data reported elsewhere in the literature, the Woschni correlation without
swirl was found to most likely under-predict the convective heat transfer by approximately
10-15%. The simulated in-cylinder pressure development for each fuel was then compared
144 Combustion Modelling and Analysis

with that of the representative non-autoigniting pressure traces acquired during octane
rating. Good agreement between the two was attained for all fuels, which suggested the
modelling approach was reasonable.

The results obtained from the combustion model were then analysed briefly. In the
absence of autoignition, the modelling results indicated that all fuels exhibited quite
similar flame propagation during octane rating. The inferred fuel mass that participated
in autoignition was also found to correlate inversely with the RON, as others have found.
Finally, the simulated unburned zone temperature histories were examined. For most
fuels, the simulated peak pressure location did not correspond to the crank angle at which
autoignition was observed during octane rating. This is thought to be in part the result of
pre-flame kinetic activity within the unburned zone, as discussed in the next chapter.
Chapter 6

Modelling Autoignition during Octane


Rating using Chemical Kinetics

6.1 Introduction

This chapter makes use of the empirically calibrated, two-zone model of the CFR engine
presented in the preceding chapter. This model is revised to simulate autoignition through
the addition of detailed chemistry within the unburned zone, but with the other sub-
models of the non-kinetic phenomena retained. In particular, the flame propagation and
in-cylinder heat transfer is modelled based on the in-cylinder pressure data measured
at standard test conditions with autoignition suppressed. The emergence of pre-flame
kinetic activity within the model, along with autoignition (or otherwise), is governed by
the detailed kinetic mechanism.

The engine model is first used to systematically investigate the effects of heat losses,
humidity, carbon monoxide and partially reacted hydrocarbons on autoignition. The
effect of nitric oxide on autoignition is then examined using the measured engine-out
emissions data presented in the preceding chapter. This analysis is used to establish
which of these factors most strongly influences pre-flame kinetic activity, the rate of fuel
oxidation and the autoignition behaviour of the various fuels. The mechanisms that
appear to be responsible for this behaviour are then compared with data reported in the
literature, in order to determine whether the modelled trends are reasonable. Finally, the
simulated in-cylinder pressure histories are compared with those measured at standard
test conditions where autoignition was observed. This comparison is used to establish

145
146 Modelling Autoignition during Octane Rating using Chemical Kinetics

whether the two-zone combustion model and detailed chemical kinetic mechanism can
reproduce the autoignition behaviour observed for the various fuels during octane rating.

6.2 Two-zone engine model with autoignition

The engine model used to study autoignition in this chapter is based on the two-zone
model of the CFR engine developed in the previous chapter. This model was revised
to simulate autoignition through the addition of a detailed chemical kinetic mechanism
within the unburned zone (Figure 6.1). However, each of the sub-models describing the
non-kinetic phenomena were retained. As such, the flame propagation and in-cylinder
heat transfer is modelled based on the in-cylinder pressure data measured at standard test
conditions with autoignition suppressed.

Figure 6.1: Schematic diagram of the revised two-zone combustion model with detailed
chemical kinetics.
6.2 Two-zone engine model with autoignition 147

6.2.1 Model formulation

As discussed in the previous chapter, the full engine model was able to simulate both the
open and closed portions of the cycle. The modelled engine operation during the open
portion of the cycle is based on the validated intake and exhaust sub-models described
in the previous chapter. These sub-models are first used to determine the volumetric
efficiency of the engine for each operating condition. The sub-models are then used in
conjunction with a residual gas sub-model to simulate the mixture state at intake valve
closure.

The modelled engine operation during the closed portion of the cycle is described in
Figure 6.2. This sub-routine includes additional calculations that account for pre-flame
kinetic activity. The mixture is first compressed by the motion of the piston during
the compression stroke. During this process, pre-flame kinetic activity that leads to
consumption of the fuel within the unburned zone is permitted. Once combustion has
commenced, consumption of the fuel may occur as the result of both pre-flame kinetic
activity within the unburned zone, along with mass transport from the unburned to the
burned zone via the flame front. The flame propagation is described using the calibrated
sub-models for the various fuels presented in the previous chapter. These sub-models
are based on the representative in-cylinder pressure traces with autoignition suppressed.
The flame is assumed to propagate spherically, with its centre located at the side-mounted
spark plug.

Since autoignition is of primary interest in this chapter, all post-autoignition combus-


tion was neglected. Therefore, once autoignition is detected by the code, the remaining
fuel was burned on the subsequent iteration. This caused a rapid increase in the simu-
lated pressure history. The standard GT-Power sub-models are then used to complete the
expansion and exhaust strokes.

Prescribing the flame propagation

As previously discussed, the flame propagation is determined by calibrating the combus-


tion model such that the simulated in-cylinder pressure history matches the representative
148 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.2: Flow chart describing the operating principle within the closed portion of the
cycle in the engine model with chemical kinetics.
6.2 Two-zone engine model with autoignition 149

pressure trace for each fuel with autoignition suppressed. This determines the amount of
mass transferred between the unburned and burned zones at each time step in the absence
of autoignition. All deviations from the prescribed flame propagation therefore occur as
the result of pre-flame kinetic activity within the unburned zone, or due to autoignition
itself.

In-cylinder heat transfer

The in-cylinder heat transfer was simulated using the Woschni correlation without swirl
(Woschni, 1967). Unless otherwise stated, all simulations were performed using the
calibrated in-cylinder heat transfer sub-models presented in the previous chapter.

Residual combustion products sub-model

The residual gas fraction is comprised of the combustion products that remain after closure
of the exhaust valve. This parameter is generally not measured, and therefore is often
estimated (Sinnamon and Sellnau, 2008; Colin et al., 2007; Cavina et al., 2004; Cho et al.,
2001). In this work, the residual gas fraction was estimated using the full engine model
presented in the previous chapter. These simulations indicated that the fraction of the
cylinder mass occupied by the residual combustion products at intake valve closure was
primarily a function of the compression ratio, as shown in Figure 6.3. This behaviour
may be expected, since the CFR engine operates unthrottled and has fixed valve timing
for all operating conditions. The estimated values, which typically range between 5-8 (%
mass), also show good agreement with equivalent data reported elsewhere in the literature
for the CFR engine (Tzanetakis et al., 2010; Alkidas, 1999; Nishiwaki, 1998; Blin-Simiand
et al., 1993; Galliot et al., 1990; Taylor, 1985; Rauckis and McLean, 1979). In all cases, the
residual products were assumed to be the ideal products of combustion. Additionally, the
measured nitric oxide (NO) and nitrogen dioxide (NO2 ) concentrations were also included
in some test cases.
150 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.3: Estimated residual gas fraction versus compression ratio. Data is presented for
the Research method SKI test condition with autoignition suppressed.

Neglecting dissociation, the combustion of any hydrocarbon fuel, Cx Hy , in the presence


of air can be expressed as,

Cx Hy + a(O2 + 3.76 N2 ) → b CO2 + c CO + d H2 O + e H2 + f O2 + 3.76 a N2 . (6.1)

The procedure used to calculate the coefficients a to f, and therefore the individual species
concentrations, is described in Appendix C. The concentration, χi, IVC , of species i within
the cylinder at intake valve closure is evaluated using the estimated fraction of the cylinder
occupied by the residual combustion products,

 
Vrf, IVC
χi, IVC = χi, products , (6.2)
Vcyl, IVC

where χi, products is the concentration of species i in the combustion products, Vrf, IVC is the
estimated volume of the cylinder occupied by the residual combustion products at intake
valve closure, and Vcyl, IVC is the total cylinder volume at intake valve closure. For CO and
NO, the engine-out exhaust exhaust emissions were measured (Table 6.1).
6.2 Two-zone engine model with autoignition 151

The blending of the fresh air-fuel mixture with the residual combustion products
is performed using the standard GT-Power sub-model. This sub-model provides the
in-cylinder concentrations of all species at intake valve closure, together with the mixture
temperature. The in-cylinder pressure at intake valve closure is estimated from the
representative pressure trace with autoignition suppressed. This information is used to
initialise the sub-routine used in the closed portion of the cycle (Figure 6.2).

Table 6.1: Estimated residual concentrations of carbon monoxide (CO) and nitric oxide
(NO) at intake valve closure. Data is presented for the Research method SKI test condition.

Fuel composition (% mol) In-cylinder concentration at IVC


Res. frac. CO NO
propane propylene n-butane iso-butane (% mass) (% mol) (ppm)
100.0 - - - 5.60 0.10 89.0
- 100.0 - - 6.60 0.01 162.0
- - 100.0 - 7.10 0.28 60.0
- - - 100.0 6.40 0.25 60.0
70.0 10.0 10.0 10.0 5.80 0.12 82.0
66.6 16.7 - 16.7 5.80 0.10 92.0
66.6 - 16.7 16.7 5.80 0.16 72.0
50.0 50.0 - - 6.10 0.07 101.0
50.0 - 50.0 - 6.50 0.23 63.0
41.7 16.6 41.7 - 6.40 0.21 67.0
16.7 - 66.6 16.7 6.90 0.27 62.0
- - 50.0 50.0 6.80 0.26 63.0

6.2.2 Chemical kinetic mechanisms

All combustion modelling presented in this chapter was undertaken using the detailed
C0 -C4 chemical kinetic mechanism developed by Healy and co-workers (Healy et al.,
2010a,b,d,c). This model contains 230 species that participate in 1319 elementary reactions.
The model has been validated over a wide range of conditions, including many of those
that are relevant to spark-ignition engines (Table 2.8). This model also appears to be the
most recent and comprehensive published in the open literature that is suitable for up to
C4 hydrocarbons.

The effect of nitrogen oxides on autoignition was examined by coupling the C0 -C4
152 Modelling Autoignition during Octane Rating using Chemical Kinetics

oxidation mechanism with a second mechanism that described the chemical pathways
by which short-chain hydrocarbons interact with NOx (Dagaut and Dayma, 2005). This
model has been validated for binary methane/ethane mixtures over a more limited range
of conditions (1-10 atm, and 800-1350 K), and therefore only contains hydrocarbon-NOx
specific reactions for methane and ethane. For longer-chain hydrocarbons, the NOx
coupling is described only through some reactions between the intermediate species
and the NOx pathways contained in the model. The combined mechanism contains 254
chemical species that participate in 1575 elementary reactions.

Coupling the engine model with detailed chemistry

The chemical kinetic mechanism was implemented in the engine model using the standard
gas-phase chemistry file. This file contained the elementary reactions and rate constant
data in CHEMKIN-format (Reaction Design, 2012). The corresponding thermodynamic
data was also imported into GT-Power from the standard CHEMKIN-format file. These
thermodynamic properties included the specific heat, enthalpy and entropy. Each property
is evaluated as a function of temperature using a 14 term polynomial fit of the data (seven
for the high temperature range, and seven for the low temperature range). All gases are
assumed to be ideal.

The coupled thermodynamic and chemical kinetic equations were integrated using
the Fortran RADAU5 sub-routine (Hairer and Wanner, 2002). This solver is based on an
implicit 5th order Runge-Kutta method.

Simulated ignition delay for the combined mechanism

The predictive capability of the combined mechanism was compared with that of the
standalone C0 -C4 oxidation mechanism using the CHEMKIN-PRO simulation tool. A
Perfectly-stirred reactor (PSR) model was used to compute the constant-volume, adiabatic
ignition delay for each of the four pure constituents, using both the standalone and
combined mechanisms. The simulations were performed for initial temperatures ranging
6.2 Two-zone engine model with autoignition 153

from 600-1050 K, and an initial pressure of 40 bar. Based on the combustion modelling
undertaken in the previous chapter, these conditions are considered representative of
those encountered in spark-ignition engines.

The mole fractions of the reactants were specified such that the relative air-fuel ratio
corresponded to the measured lambda values for Standard Knock Intensity (SKI) in the
engine experiments (Table 5.1). The initial mole fractions of the oxides of nitrogen were
specified as zero. Therefore, this analysis directly evaluates the influence of the additional
reactions in the NOx mechanism on the simulated ignition delay. The ignition delay was
evaluated based on the elapsed time for the reactor to reach a temperature of 200 K above
its initial value (Konnov, 2008).

The ratio of the ignition delay obtained using the standalone mechanism to that
obtained using the combined mechanism is presented in Figure 6.4. It is evident that
the inclusion of the NOx sub-model has a negligible effect on the simulated ignition
delay. This behaviour is observed for each of the pure constituents, over the full range of
temperatures considered. This combined mechanism is utilised in all of the subsequent
modelling presented in this chapter.

Figure 6.4: Ratio of the constant-volume ignition delay obtained using the standalone
mechanism to that obtained using the combined mechanism. All simulations were per-
formed with an initial pressure of 40 bar.
154 Modelling Autoignition during Octane Rating using Chemical Kinetics

6.3 Effect of heat losses on autoignition

The full engine model with chemical kinetics is first used to examine the sensitivity of
propane autoignition to heat losses. The value of the Woschni correlation convective
heat transfer coefficient multiplier, ζ, was varied between 0.50 and 1.50. This variation
represents a 50% increase/decrease in the computed convective heat transfer coefficient
obtained from the unmodified Woschni correlation without swirl (Equation 5.4). All other
engine operating variables within the model corresponded to the Research method SKI
test condition for this fuel.

Figure 6.5 indicates that the value of ζ strongly influences the autoignition onset in
a physically intuitive manner. In the case of reducing ζ, the reduced convective heat
transfer between the working fluid and the combustion chamber surfaces causes the
unburned zone temperature to increase significantly. This causes autoignition to occur
more prematurely, due to the strong temperature dependence of the chemical reactions.
The same simulations performed for the other fuels considered in this work produced
similar trends, indicating that heat transfer is a significant contributor to autoignition.

6.4 Effect of humidity on autoignition

In the Research and Motor methods, the intake air humidity must be within the prescribed
limits shown in Table 6.2. This requirement is satisfied by firstly chilling the intake air to a
temperature of 2 ◦ C using a dehumidifying unit, and subsequently re-heating the air prior
to entering the engine (refer to Section 3.1).

To examine the sensitivity of the autoignition onset to humidity, simulations were


performed with varying amounts of water added to the intake air. Humidity effects can be
attributed to the higher heat capacity of H2 O relative to that of air. This thermodynamic
effect reduces the compressed mixture temperature with increasing H2 O concentration,
and therefore postpones autoignition (Dec et al., 2007; Goy et al., 2001; Lanzafame, 1999).
These simulations were compared with one that assumed the air entering the engine was
6.4 Effect of humidity on autoignition 155

Figure 6.5: Effect of variations in the Woschni convective heat transfer coefficient multiplier
on the simulated autoignition of propane. Positive heat transfer rates indicate that heat
is transferred from the working fluid to the combustion chamber surfaces. Simulated
conditions correspond to the Research method SKI test condition.
156 Modelling Autoignition during Octane Rating using Chemical Kinetics

Table 6.2: Intake air humidity limits prescribed in Research and Motor methods (ASTM
International, 2012a,b).

Intake air humidity


(kg water/ kg dry air)

Lower limit 0.00356


Upper limit 0.00712

completely dry (Figure 6.6). These results indicate that the autoignition onset does not
vary significantly with the humidity in compliant ASTM testing. Further, the assumption
that dry air enters the engine therefore appears to be reasonable, and is used in all of the
subsequent modelling presented in this chapter.

6.5 Effect of carbon monoxide (CO) on autoignition

The presence of CO within the residual combustion products can primarily be attributed
to the incomplete combustion of a fuel in the presence of less than the stoichiometric
air requirement (Heywood, 1988). The effect of CO on autoignition was examined by
systematically varying the CO concentration at intake valve closure between the measured
residual quantity present in the engine, and eight times this value (Table 6.1). The latter
far exceeds the concentration that could be present within the engine, but provides an
insight into the sensitivity of autoignition to CO.

Figure 6.7 indicates that no strong effect on autoignition can be associated with CO. In
most cases, the presence of elevated CO concentrations does not begin to influence the
autoignition behaviour significantly until its residual concentration exceeds approximately
four times the measured residual quantity. For both n-butane and iso-butane, autoignition
was no longer observed once the CO concentration exceeded four times the measured
value.

The oxidation of CO primarily occurs via the pathway described by Reaction R 7.1.
However, this reaction occurs much more slowly than the rate of initial CO formation
6.5 Effect of carbon monoxide (CO) on autoignition 157

Figure 6.6: Effect of intake air humidity on autoignition. Simulated conditions correspond
to the Research method SKI test condition.
158 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.7: Simulated variation in the autoignition onset for the pure constituents with the
addition of CO. Filled icons represent the inferred residual CO concentrations at intake
valve closure. Simulated conditions correspond to the Research method SKI test condition.

during the oxidation of the fuel (Hellier, 2013).

CO + OH → CO2 + H
(R 7.1)

It also also possible for the oxidation of CO to occur via the hydroperoxyl radical (Reaction
R 7.2). However, this pathway is generally less well favoured, due to the even slower
reaction rate in comparison with R 7.1 (Bowman, 1975).

CO + HO2 → CO2 + OH
(R 7.2)

Since Reactions R 7.1 and R 7.2 occur much more slowly than the reactions between the
6.6 Effect of unreacted and partially reacted hydrocarbons on autoignition 159

intermediate species and the radicals, the observed inhibiting behaviour with increasing
CO concentration is physically intuitive. This is also consistent with the high RON of
CO, which exceeds the upper 120.3 limit of the ON scale (Puckett, 1945). This value far
exceeds that of all the gaseous fuels examined in this work, and therefore the addition of
CO inhibits the onset of autoignition. Collectively, these trends are also consistent with
those reported elsewhere in the literature (Machrafi et al., 2008; Dubreuil et al., 2006).

6.6 Effect of unreacted and partially reacted hydrocarbons on au-


toignition

Unreacted and partially reacted hydrocarbons may originate as the result of several
complex processes. These often include the desorption of hydrocarbon-rich gases from
within the oil layers and crevice volumes where the flame is unable to propagate, together
with incomplete combustion itself. Given the complexity and cycle-to-cycle variations
associated with these processes, the residual combustion products are comprised of a
wide range of species. In most cases, these hydrocarbons are not individually identified in
standard engine tests.

The effect of unreacted and partially reacted hydrocarbons on the autoignition of


propane was examined using the speciated exhaust emissions data reported by Kaiser
et al., 1983. This data was acquired using a flame-ionization detector, and contained
the exhaust concentrations of the unreacted propane, together with a range of stable
intermediate species (Figure 6.8). This analysis technique was reported to have identified
more than 90% of the total hydrocarbons in the exhaust (Appendix D).

The raw speciated data reported by Kaiser et al. was used to compute the residual
concentrations of the same hydrocarbons that would be expected to be present for propane
at the Research method SKI test condition (Table 6.3). These calculations assume the
total engine-out hydrocarbon emissions are 1500 ppm C1 (Appendix D), as measured in
the engine experiments. Further, it was assumed that the residual combustion products
occupied 5.60 (% mass) of the cylinder at intake valve closure, as previously estimated
(Table 6.1). This corresponds to a residual fraction of 5.87 (% vol), and therefore a total
160 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.8: Engine-out hydrocarbon emissions from a propane-fuelled spark-ignition


engine (adapted from Kaiser et al., 1983).

residual HC concentration of 88.1 (ppm C1 equivalent) at intake valve closure.

Table 6.3: Calculated residual hydrocarbon concentrations at intake valve closure, based
on the data reported for propane (Kaiser et al., 1983).

Concentration at IVC
Hydrocarbon (ppm) (ppm C1 equiv.)
propane (C3 H8 ) 17.5 52.5
propylene (C3 H6 ) 2.7 8.1
ethane (C2 H6 ) 0.8 1.6
ethylene (C2 H4 ) 7.0 14.0
acetylene (C2 H2 ) 3.1 6.2
methane (CH4 ) 5.7 5.7

88.1

The effect of these species on autoignition was examined with and without NO in-
cluded in the model. Figure 6.9 indicates that the simulated trends for the two test cases
differ considerably. In the first case where NO was neglected, these hydrocarbons have a
minor promoting effect on autoignition with increasing concentration (albeit well outside
the practical limits for these species in the engine). In the second case, where the residual
6.6 Effect of unreacted and partially reacted hydrocarbons on autoignition 161

Figure 6.9: Effect of unreacted and partially reacted hydrocarbons on the simulated
autoignition onset of propane. The composition and relative proportions of these hydro-
carbons were modelled using the data reported by Kaiser et al. (Figure 6.8). Simulated
conditions correspond to the Research method SKI test condition.
162 Modelling Autoignition during Octane Rating using Chemical Kinetics

NO emissions were included in the model, this promoting effect on autoignition is not
observed. Instead, all test cases exhibit near identical autoignition onset locations. Con-
sistent with the literature, this indicates that the promoting effect of NO on autoignition
is stronger than that of much larger concentrations of unreacted and partially reacted
hydrocarbons. For example, Chan et al. reported that the addition of 3000 ppm of ethane
had the equivalent promoting effect on methane oxidation as just 10 ppm of NO (Chan
et al., 2012).

6.7 Effect of nitrogen oxides (NOx ) on autoignition

6.7.1 Constant-volume ignition delay

The isolated effect of NO on autoignition was first examined using a constant-volume,


PSR model. This technique avoids the complexity of the detailed engine model, where
variations in flame propagation, residual combustion products and heat transfer occur
between the various test cases. The PSR model therefore allows the autoignition behaviour
to be compared over a range of reference conditions common to each fuel.

The constant-volume ignition delay for the four pure constituents was evaluated
for initial temperatures ranging between 600-1050 K, and an initial pressure of 40 bar.
Based on the combustion modelling undertaken in the previous chapter, these conditions
are considered representative of those encountered in spark-ignition engines. The NO
concentration within the reactor was also varied between 0-400 ppm. The ignition delay
was evaluated based on the elapsed time for the reactor to reach a temperature of 200 K
above its initial value (Konnov, 2008).

Consistent with the trends observed in the previous section for propane, Figure 6.10
indicates that the autoignition behaviour of all fuels is influenced by NO concentrations
of as little as 50 ppm. For the three paraffins, the ignition delay in the presence of all
NO concentrations is shorter than that of the neat fuel. Propylene, however, exhibits
contrasting behaviour, with the ignition delay in the low and intermediate temperature
6.7 Effect of nitrogen oxides (NOx ) on autoignition

Figure 6.10: Simulated constant-volume autoignition delays for the pure constituents, with varying amounts of NO. All simulations
were performed with an initial pressure of 40 bar, with lambda corresponding to the Research method SKI test condition for each
163

fuel.
164 Modelling Autoignition during Octane Rating using Chemical Kinetics

regimes monotonically increasing with the addition of NO. This trend disagrees with the
limited data reported for propylene, which indicates this fuel is generally an effective
promoter of NO to NO2 conversion (Shin and Yoon, 2003; Hori et al., 1998; Penetrante et al.,
1997; Bromly et al., 1992; Bromly, 1991). This behaviour would be expected to promote
the oxidation of propylene in the presence of NO, and therefore reduce its ignition delay.
Given the apparent significance of NO on the autoignition characteristics of these fuels,
this compound is further examined in the subsequent sections of the chapter using the
full engine model.

6.7.2 Effect of nitric oxide (NO) on autoignition

The effect of NO on autoignition was first examined for the four pure constituents. The
results obtained from these simulations are compared with the measured in-cylinder
pressure histories in Figure 6.11.

Consistent with the trends observed in the constant-volume ignition delay modelling,
the promoting effect of NO on autoignition is modelled across all test cases other than
propylene. For the three paraffins, the simulated autoignition location with NO included
in the model occurs between 1.0 and 3.0 crank angle degrees (0.28 ms to 0.83 ms) more
prematurely, and shows reasonable agreement with experiment. This suggests that the
model is capable of reproducing the observed autoignition behaviour of these hydrocar-
bons during octane rating. Importantly, the promoting effect of NO on the autoignition
of these hydrocarbons is also consistent with data reported elsewhere in the literature
(Bromly, 1991; Bromly et al., 1992; Hori et al., 1998; Dagaut et al., 2001; Nelson and Haynes,
1994).

As expected from the previous analysis, the simulated autoignition behaviour of


propylene differs to the three paraffins. Figure 6.11 indicates that the inclusion of NO in
the model has a dramatic inhibiting effect on the autoignition onset of approximately 10.0
crank angle degrees (2.78 ms). Consequently, autoignition occurs outside of the prescribed
burn period for this fuel, and is therefore non-physical. Equivalent simulations performed
for varying NO concentrations indicate this trend is monotonic in nature, and increases
6.7 Effect of nitrogen oxides (NOx ) on autoignition 165

Figure 6.11: Measured and simulated pressure histories for the pure constituents, with
and without NO. Data is presented for the Research method SKI test condition.
166 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.11: Measured and simulated pressure histories for the pure constituents, with
and without NO. Data is presented for the Research method SKI test condition.
6.7 Effect of nitrogen oxides (NOx ) on autoignition 167

Figure 6.12: Simulated variation in the autoignition onset for the pure constituents with
the addition of NO. Simulated conditions correspond to the Research method SKI test
condition.

rapidly at approximately 120 ppm (Figure 6.12). Further, for NO concentrations exceeding
approximately 200 ppm at intake valve closure (approximately 3100 ppm within the
exhaust), autoignition was fully suppressed.

This poor agreement for propylene can likely be attributed to a combination of several
factors. First, the mechanism does not contain fuel-specific hydrocarbon-NOx reaction
pathways for any hydrocarbons larger than ethane (refer to Section 6.2.2). Therefore, the
chemical interactions between larger hydrocarbons and NOx are only via the reactions
common to methane and ethane. The results presented in Figure 6.11 may therefore
indicate that specific propylene-NOx interactions are more important than the equivalent
reactions for the three paraffins, and that the absence of these reactions leads to a net
inhibiting effect on autoignition.

Propylene also has the highest residual NO concentration of all the fuels considered in
this work, owing to its lean test condition during octane rating (Table 6.1). Therefore, these
results could also indicate an underlying sensitivity within the mechanism to high residual
168 Modelling Autoignition during Octane Rating using Chemical Kinetics

NO concentrations. Given the trends exhibited by the three paraffins in Figure 6.12, this
factor is considered less likely, but is nonetheless explored subsequently in Section 6.9.

A further contributing factor may also include the known uncertainties in the selection
of key reaction rate constants for propylene oxidation. The effect of these uncertainties on
the ignition delay has been widely discussed, particularly for the high pressure conditions
relevant to combustion engines (Burke and Curran, 2013).

Multi-component fuels

The same simulations were performed for eight mixtures of the pure constituents. Addi-
tional simulations were also performed with the NO concentration at intake valve closure
corresponding to double the measured value reported in Table 6.1. The results obtained
from these simulations are compared with the experimental autoignition locations in
Figure 6.13.

Figure 6.13: Measured and simulated autoignition onset for Propane (Pr), Propylene (Pryl),
n-Butane (nB), iso-Butane (iB) and mixtures thereof. Autoignition was not observed for
propylene when the residual NO concentration was doubled. Data is presented for the
Research method SKI test condition.
6.7 Effect of nitrogen oxides (NOx ) on autoignition 169

Consistent with the previous data, the test case with NO included in the model shows
good agreement with the experimental autoignition onset for most fuels. Excluding
propylene, the standard error (SE) between the simulation and experimental data is
approximately 1.3 crank angle degrees (0.36 ms). In most cases, the simulated autoignition
onset occurs marginally earlier than observed in the engine experiments. The exception to
this behaviour is iso-butane, and the two mixtures with significant n-butane/iso-butane
content. Importantly, the autoignition location observed in the engine experiments is
bracketed by the two most extreme simulated test cases. This provides further evidence
which suggests the model is capable of reproducing the autoignition behaviour of the
paraffins.

6.7.3 Fuel oxidation in the presence of nitric oxide

The simulated oxidation of the various fuels in the lead up to autoignition is presented
in Figure 6.14. Consistent with the previous results, the oxidation of each fuel (with the
exception of propylene) proceeds more quickly in the presence of NO. However, for most
fuels, there is a minor difference between the three tests cases until approximately 50% of
the fuel has been oxidised. The exception to this behaviour appears to be the fuels with
significant n-butane content.

The more premature oxidation of n-butane in the presence of NO is likely to be


associated with the increased yields of key intermediate species at an earlier stage in the
combustion event. Figure 6.15 shows that the concentration of hydrogen peroxide (H2 O2 ),
for example, is approximately an order of magnitude higher at the TDC location (0 crank
angle degrees) in the presence of NO. The decomposition of H2 O2 produces highly reactive
hydroxyl radicals via Reaction R 7.3, which readily promotes fuel oxidation. Under many
conditions, this is sufficient for autoignition to occur (Bahrini et al., 2012; Beerer and
McDonell, 2011; Smith et al., 1985; Downs et al., 1951).

H2 O2 (+M) → OH + OH (+M) (R 7.3)


170 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.14: Simulated fraction of the starting fuel oxidised versus crank angle for Propane
(Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB) and mixtures thereof. Simulated
conditions correspond to the Research method SKI test condition.
6.7 Effect of nitrogen oxides (NOx ) on autoignition 171

Figure 6.14: Simulated fraction of the starting fuel oxidised versus crank angle for Propane
(Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB) and mixtures thereof. Simulated
conditions correspond to the Research method SKI test condition.
172 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.15: Simulated species concentration histories for n-butane oxidation versus crank
angle. Simulated conditions correspond to the Research method SKI test condition.

Figure 6.14 also indicates that the oxidation of iso-butane is less sensitive to NO
than its straight-chain isomer. Iso-butane is more branched than n-butane, and therefore
has a higher percentage of primary hydrogen atoms. These atoms have higher bond
dissociation energies than secondary and tertiary sites, and therefore the oxidation of
iso-butane initially produces lower yields of HO2 than n-butane. Consequently, the
mutual sensitisation of the parent hydrocarbon and NO, which may occur via the pathway
described by Reaction R 7.4, should be less well favoured relative to n-butane.

NO + HO2 → NO2 + OH (R 7.4)

There is evidence of this behaviour in Figure 6.16, which indicates that the production
of HO2 during the oxidation of iso-butane is very low. The peak concentration of HO2 , for
example, is less than one-third of that observed during the oxidation of n-butane. This is
despite the higher pressures encountered during standard octane rating conditions for
iso-butane (owing to its higher RON, and therefore higher compression ratio). Under
6.7 Effect of nitrogen oxides (NOx ) on autoignition 173

Figure 6.16: Simulated species concentration histories for iso-butane oxidation versus
crank angle. Simulated conditions correspond to the Research method SKI test condition.

otherwise equivalent operating conditions, these higher pressures are likely to favour the
production of additional HO2 (Moreac, 2003).

The simulation results for iso-butane also indicate that the H2 O2 concentration at any
given instant does not differ significantly when NO is included in the model. Since the
autoignition of iso-butane under engine-relevant conditions is primarily a high tempera-
ture process that occurs at approximately 950 K (Green et al., 1987b), the small variations
in H2 O2 production with NO addition may explain the lower sensitivity of iso-butane
to NO than observed for the other fuels. This high temperature autoignition process is
dominated by the chain branching sequence described by Reaction R 7.5 and subsequently
by Reaction R 7.3,

iso − C4 H10 + HO2 → C4 H9 + H2 O2


(R 7.5)
174 Modelling Autoignition during Octane Rating using Chemical Kinetics

where C4 H9 can refer to either the iso-butyl or t-butyl radical. Therefore, the minor
variations observed in the H2 O2 concentrations between the two test cases is likely to be a
contributor to the lower sensitivity exhibited by iso-butane to NO.

Consistent with the previous results, the simulated species histories for propylene also
differ considerably to those observed for the paraffins. Figure 6.17 indicates that both
H2 O2 and HO2 are readily formed in the absence of NO. However, the inclusion of NO
in the model inhibits the formation of these species until approximately 15.0 crank angle
degrees. This behaviour again disagrees with the limited experimental data for propylene
(Hori et al., 1998; Bromly, 1991).

Figure 6.17: Simulated species concentration histories for propylene oxidation versus
crank angle. Simulated conditions correspond to the Research method SKI test condition.

6.7.4 Effect of nitrogen dioxide (NO2 ) on autoignition

NO is the dominant oxide of nitrogen originating from a spark-ignition engine operated


in the vicinity of the stoichiometric condition, and typically accounts for at least 95% of
6.7 Effect of nitrogen oxides (NOx ) on autoignition 175

the total engine-out NOx emissions (Heywood, 1988; Hilliard and Wheeler, 1979; Hanson
and Egerton, 1937). The remaining 5% is typically comprised of nitrogen dioxide (NO2 ).
Therefore, the concentration of NO2 within the fresh mixture at intake valve closure is
likely to be quite small. Despite this, several studies have shown that small amounts of
NO2 often influence the reactivity of a system (Chan et al., 2011; Bendtsen et al., 2000).

The engine-out emissions of NO2 were unable to be measured during the engine
experiments. The effect of NO2 was instead examined by assuming its contribution
to the total engine-out NOx emissions was 5%. This assumption yields residual NO2
concentrations at intake valve closure of approximately 3-8 ppm. The results obtained
from these simulations are compared with the previous simulation results, along with
the experimental data in Figure 6.18. These results indicate that the inclusion of small
amounts of NO2 does not have a significant influence on the simulated autoignition
location, beyond that encountered with NO already included in the model.

Figure 6.18: Simulated autoignition location when the total engine-out NOx emissions
were assumed to be comprised of 95% NO and 5% NO2 . Data is presented for Propane
(Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB) and mixtures thereof, at the Research
method SKI test condition.
176 Modelling Autoignition during Octane Rating using Chemical Kinetics

6.8 Pre-flame kinetic activity within the unburned zone

6.8.1 Fuel consumption

The fraction of the starting fuel consumed by pre-flame kinetic activity within the un-
burned zone versus crank angle is presented for the four pure constituents in Figure 6.19.
This property is inclusive of the fuel initially consumed during the compression of the
mixture by both the piston and the flame, along with the fuel subsequently consumed by
autoignition. The latter can be identified as the near instantaneous consumption of the
remaining fuel at approximately 10.0 crank angle degrees (20.0 crank angle degrees for
propylene in the presence of NO).

For each of the paraffins, the fuel consumed during the compression of the unburned
zone by the flame is higher in the presence of NO. Similarly, a greater fraction of the
starting fuel is consumed by autoignition when NO is included in the model (more than
30% and 40% for propane and n-butane, respectively), with autoignition occurring more
prematurely than observed when NO was neglected. In the case of n-butane, a larger
fraction of the starting fuel is consumed during the compression of the unburned zone
than for the other fuels. This is consistent with motored engine studies for this fuel
(Addagarla, 1991; Green et al., 1987a). The two-stage autoignition behaviour of n-butane
is also apparent.

6.8.2 Unburned zone temperature increase prior to autoignition

It was identified in the previous chapter that the effects of pre-flame kinetic activity arising
from compression of the unburned zone by the piston and the flame may be significant
for the fuels considered in this work. This is now examined. Figure 6.20 presents the
temperature difference between the simulated unburned zone temperatures presented in
the previous chapter using equilibrium calculations, and those simulated using the engine
model with chemical kinetics. Since the heat transfer, flame propagation and residual gas
sub-models are equivalent for each fuel, the variations in the unburned zone temperature
6.8 Pre-flame kinetic activity within the unburned zone 177

Figure 6.19: Simulated fraction of the starting fuel consumed by pre-flame kinetic activity
within the unburned zone versus crank angle. Data is presented for the Research method
SKI test condition.
178 Modelling Autoignition during Octane Rating using Chemical Kinetics

can be attributed to heat release from pre-flame kinetic activity.

It is evident in Figure 6.20 that pre-flame kinetic activity influences the temperature of
the mixture even prior to the start of combustion. In all cases, the mixture temperature at
spark discharge is approximately 20 K higher than the value obtained from the equivalent
chemical equilibrium simulations presented in the previous chapter. This trend is largely
independent of both fuel composition and the small variations in the compression ratio
that occur from fuel-to-fuel. The inclusion of NO in the model also further influences the
evolution of the unburned zone temperature. The latter increases more rapidly in the
presence of NO for all fuels, with the exception of propylene, and coincides with a more
premature autoignition onset location. Collectively, these results suggest that pre-flame
kinetic activity is a significant contributor towards the autoignition of the fuels examined
in this work.

These simulation results also indicate that the fuels with higher RONs generally tend
to exhibit a marginally smaller degree of temperature increase (due to pre-flame kinetic
activity) prior to autoignition. This trend is presented in Figure 6.21. In the case of propane
and n-butane, these observations are consistent with end-gas temperature data acquired
using Coherent anti-Stokes Raman spectroscopy (CARS) (Kalghatgi et al., 1995b,a; Bradley
et al., 1994; Lucht et al., 1987). Of further note is that the inclusion of NO does not appear
to significantly alter the unburned gas temperature at which autoignition occurs, with
the exception of propylene. This observation is fundamentally the same as that reported
elsewhere in the literature (Green et al., 1987b; Gluckstein and Walcutt, 1961), and suggests
that autoignition will occur for each fuel once its critical condition has been reached. The
presence of NO does not alter this temperature, but only reduces the time taken to reach
this condition.

6.9 Further analysis of propylene autoignition

The modelling presented in the previous sections of this chapter has mostly shown
good agreement with the engine experiments. The exception to this behaviour has been
6.9 Further analysis of propylene autoignition 179

Figure 6.20: Unburned gas temperature increase above the temperature obtained from
the combustion simulation with chemical equilibrium in Chapter 5. Data is presented for
Propane (Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB) and mixtures thereof, at the
Research method SKI test condition.
180 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.20: Unburned gas temperature increase above the temperature obtained from
the combustion simulation with chemical equilibrium in Chapter 5. Data is presented for
Propane (Pr), Propylene (Pryl), n-Butane (nB), iso-Butane (iB) and mixtures thereof, at the
Research method SKI test condition.
6.9 Further analysis of propylene autoignition 181

Figure 6.21: Temperature rise within the unburned zone prior to autoignition arising from
pre-flame kinetic activity versus RON.

propylene. For this fuel, the standard octane tests are conducted under lean mixture
conditions (λ = 1.03). This differs from all other fuels considered, for which the SKI test
condition occurred with rich mixture conditions. These differences have also been shown
to cause large variations in the residual NO concentrations at intake valve closure.

To evaluate the effects of these variations on the autoignition behaviour, simulations


were performed for both propane and propylene, with the mixture condition for each
varied from that encountered in the standard octane tests. In-cylinder pressure data was
acquired for propane (λSKI =0.96) under rich (λ=0.90) and lean (λ=1.03) operating condi-
tions. Similarly, data was acquired for propylene (λSKI =1.03) under stoichiometric and rich
(λ=0.96) operating conditions. Non-autoigniting pressure data was unable to be acquired
for propylene at mixture conditions richer than λ=0.96, due to the reduced effectiveness of
the dilute TEL in suppressing autoignition. All other operating conditions corresponded
to those at the SKI test condition (Table 3.3). The residual CO and NO concentrations at
intake valve closure were determined using the same procedure described previously in
Section 6.2.1, and are presented in Table 6.4.
182 Modelling Autoignition during Octane Rating using Chemical Kinetics

Table 6.4: In-cylinder CO and NO concentrations at intake valve closure for propane and
propylene under standard and non-standard octane rating conditions.

In-cylinder concentration at IVC

CO NO
Operating condition (% mol) (ppm)

Propane
λ = 0.90 0.20 49.0
λ = 0.96 (SKI) 0.10 89.0
λ = 1.03 0.01 130.0

Propylene
λ = 0.96 0.12 87.0
λ = 1.00 0.05 124.0
λ = 1.03 (SKI) 0.01 162.0

The measured and simulated pressure histories for propane and propylene under these
non-standard octane rating conditions are presented in Figures 6.22 and 6.23 respectively.
For propane, these results demonstrate that the simulated autoignition onset generally
agrees well with experiment under both rich and lean conditions. This is consistent with
the agreement observed previously for the SKI test condition (Figure 6.11). Intuitively,
these results also demonstrate that the inclusion of NO in the model has the greatest
influence on the simulated autoignition onset under lean conditions, where the residual
NO concentration is considerably higher (Table 6.4).

For propylene, the better agreement between the simulated autoignition onset and
experiment is similarly obtained under rich conditions, as shown in Figure 6.24. However,
as the value of lambda is progressively increased, the agreement between the simulated
autoignition location and the experimental data deteriorates. Of further note is that the
inclusion of NO in the model has an inhibiting effect on the autoignition of propylene
in all cases, which disagrees with fundamental experiments (Hori et al., 1998; Bromly,
1991). Thus, despite the satisfactory performance of the kinetic model for the paraffins,
the same agreement could not be obtained for propylene in terms of absolute accuracy,
nor the expected trend with increasing NO concentration. This suggests that a number of
key chemical reactions between propylene, NO and/or the intermediate species are not
6.9 Further analysis of propylene autoignition 183

Figure 6.22: Measured and simulated pressure histories for propane under non-standard
octane rating conditions.
184 Modelling Autoignition during Octane Rating using Chemical Kinetics

Figure 6.23: Measured and simulated pressure histories for propylene under non-standard
octane rating conditions.
6.10 Summary 185

Figure 6.24: Relative error in the simulated autoignition onset versus residual NO concen-
tration at intake valve closure.

present within the model. Thus, further work on the autoignition chemistry of propylene
in particular appears to be required.

6.10 Summary

This chapter presented a systematic modelling study of the autoignition behaviour of


twelve LPG fuels during octane rating. The work made use of the empirically calibrated,
two-zone model of the CFR engine presented in the preceding chapter, which was modified
to incorporate a detailed chemical kinetic mechanism.

The model was first used to examine the effects of heat losses, humidity, carbon
monoxide and partially reacted hydrocarbons on autoignition. When varied within
physically reasonable limits, the presence of neither carbon monoxide or the partially
reacted hydrocarbons appeared to have a significant effect on the simulated autoignition
behaviour. Similarly, humidity levels that did not exceed the upper limit prescribed in the
standard test methods for octane number had no significant effect on autoignition. Heat
186 Modelling Autoignition during Octane Rating using Chemical Kinetics

losses, however, were shown to strongly influence autoignition in a physically intuitive


manner.

The effect of small amounts of nitric oxide and nitrogen dioxide (NOx ) on autoignition
was then examined using the measured engine-out emissions data presented in the
preceding chapter. This analysis demonstrated that nitric oxide had a promoting effect on
the pre-flame kinetic activity, the rate of fuel oxidation, and the location of the autoignition
onset for all fuels, with the exception of propylene. The promoting effect of nitric oxide was
most pronounced in fuels primarily comprised of n-butane. This behaviour was attributed
to variations in the concentrations and timing of key intermediate species. However, the
presence of nitric oxide caused no significant changes to the simulated unburned zone
temperature at which autoignition was observed for any of the paraffins. Rather, nitric
oxide only reduced the time taken to reach this temperature. Nitrogen dioxide was found
to have an insignificant effect on autoignition in comparison with nitric oxide, due to its
very low concentration.

The autoignition characteristics of each fuel were then compared with the representa-
tive autoigniting pressure data acquired during octane rating. In general, good agreement
was obtained between the simulations and the experimental results for most fuels, pro-
vided that reasonable concentrations of nitric oxide was included in the model. However,
the simulation results for the fuels with significant iso-butane content agreed marginally
less well than the other paraffins. Overall, this suggested the combustion model was
largely capable of reproducing the experimental autoignition behaviour of the paraffins.

Nonetheless, the level of agreement between the simulation results and the experi-
mental data was generally quite poor for propylene. This behaviour could in part be
explained in terms of lambda variations, and thus the recirculated nitric oxide. However,
all simulation results indicated that nitric oxide had an inhibiting effect on propylene
oxidation, which disagreed with the limited experimental data reported in the literature.
Thus, further work on the autoignition chemistry of propylene in particular appears to be
required.
Chapter 7

Conclusions and recommendations for


future research

This thesis examined the octane numbers and autoignition behaviour of LPG mixtures in
a CFR engine. Since there were no active test methods or hardware that could be used
to measure the octane numbers of LPGs, it was first necessary to develop a fuel system
that was capable of preparing and dispensing gaseous fuels. This system was comprised
of four Mass Flow Controllers, and could be used to prepare up to four component LPG
mixtures comprised of propane, propylene (propene), n-butane and iso-butane. The
manner in which the gaseous fuel was delivered into the intake charge closely followed
that prescribed in the now obsolete Motor (LP) method (ASTM International, 1986), which
was formally used to measure the MON of LPG prior its withdrawal in 1989.

The RONs and MONs for a broad range of mixtures were first measured. Good
agreement was obtained with the more limited data of Boldt (Boldt, 1967), thus validating
the hardware and test procedure. The remaining 50 mixtures then more comprehensively
mapped the RON and MON space for LPG.

The autoignition behaviour of various fuels was then examined. An empirically cal-
ibrated model of the CFR engine was developed and validated using the commercial
GT-Power software tool. A technique that decoupled the kinetic and non-kinetic phenom-
ena observed at standard test conditions was used to calibrate the various sub-models,
using measured in-cylinder pressure data with autoignition suppressed. This decoupled
approach was necessary, as full simulation of turbulent combustion and other phenomena

187
188 Conclusions and recommendations for future research

that occur in an engine, together with detailed chemistry, is computationally very chal-
lenging and outside the scope of this work. The calibrated sub-models that described the
non-kinetic phenomena were then transplanted into an equivalent two-zone model that
incorporated detailed chemical kinetics. This model was able to simulate the autoignition
of the various fuels. These simulation results were compared with the in-cylinder pressure
data measured under standard octane rating conditions, which exhibited autoignition.

7.1 Conclusions

In reference to the research questions proposed in Section 2.7, the following conclusions
have been made.

1. Do the RONs and MONs of propane, propylene, n-butane and iso-butane mixtures
vary antagonistically and/or synergistically?

Based on the set of RON and MON data presented in this thesis, it appears as though the
assumption of linear blending is reasonable for practical LPG fuels. In the case of the MON,
all of the observed non-linearity was synergistic, and generally did not exceed 1.0 octane
number. For the RON, both synergistic and antagonistic blending of up to approximately
1.5 octane numbers was observed in the binary mixtures. Most of this non-linearity in the
RON could be associated with propylene, which blended synergistically with n-butane
and iso-butane, but antagonistically with propane. This generally resulted in quite linear
blending in LPGs that contained at least three constituents (i.e. practical fuels). LPGs
comprised of only paraffins always blended close to linearly in the RON for all mixtures.

These observations have practical significance, since the currently active EN 589 (Euro-
pean Committee for Standardisation, 2008) and ASTM D2598 (ASTM International, 2007a)
standards specify a linear sum calculation method for the MON of LPG. The experimental
data showed that these two calculation methods were able to predict the MON of LPG
with overall standard errors of 0.6 and 0.7 octane number respectively. In the case of
the EN 589 standard, there was also no evidence that suggested this calculation method
7.1 Conclusions 189

was unsuitable for LPGs with significant propylene content, as has been suggested (Aus-
tralian Government, Department of Environment and Heritage, 2001; Australian Liquefied
Petroleum Gas Association, 2000).

2. How do the existing LPG fuel quality standards constrain the RON of LPGs in
comparison with the minimum requirements that are directly mandated in fuel quality
standards for gasolines?

Unlike gasolines, the minimum RON of LPG is not specified in fuel quality standards,
despite this property appearing to be most relevant to modern, spark-ignition engines.
Instead, the RON of LPG is indirectly controlled using vapour pressure and MON require-
ments (e.g. Australia and Europe) or direct compositional requirements (e.g. Canada and
the United States).

Imposition of the vapour pressure and MON requirements specified in the Australian
and European LPG fuel standards indicated that these requirements impose significant
restrictions on the composition of LPG. The Australian fuel standard constrains the RON
of all LPGs comprised of propane, propylene, n-butane and iso-butane to be at least 99.5
throughout the year. The minimum RON of LPG in Europe can be more variable, due to
the seasonally varying lower vapour pressure requirement. The most stringent Grade A
vapour pressure requirement used in the winter months constrains the RON of LPG to
be at least 99.8. However, during warmer periods of the year, the RON of LPG may drop
as low as 93.5 (Grade E). Thus, in all cases, the RON of Australian and European LPG
is higher than that of standard gasolines in Australia (RON 91.0). This suggests that the
existing fuel standards unnecessarily restrict the composition of commercial LPGs, given
the widespread use of LPG in retro-fitted gasoline vehicles at present.

The compositional requirements of the HD-5/Special-duty Propane fuel standard used


in Canada and the United States constrains the minimum RON of LPG to be even higher
than that in Australia and Europe. Since this standard imposes a high propane content,
the minimum RON of LPG is 108.2.
190 Conclusions and recommendations for future research

3. What are the phenomena that need to be included in an accurate model of LPG
autoignition in a spark-ignition engine?

The contribution of various phenomena to LPG autoignition in a spark-ignition engine


was investigated by coupling the two-zone model of the CFR engine presented in Chapter
5 with the detailed chemical kinetic mechanisms developed by Healy et al., 2010a,b,d,c
and Dagaut and Dayma, 2005. The flame propagation was modelled for each fuel using
in-cylinder pressure data acquired under standard octane rating conditions, but with
autoignition suppressed.

When the residual gas concentrations of carbon monoxide and the partially reacted
hydrocarbons were varied within reasonable limits, neither of these factors appeared to
have a significant effect on the simulated autoignition behaviour. Similarly, humidity
levels that did not exceed the upper limit prescribed in the standard test methods for
octane number had no significant effect on autoignition.

However, low concentrations of nitric oxide in the residual gas was found to be a
significant contributor to engine autoignition. Nitric oxide had a promoting effect on the
pre-flame kinetic activity, the rate of fuel oxidation and the autoignition onset location
for all fuels, with the exception of propylene. Nitrogen dioxide was found to have an
insignificant effect on autoignition in comparison with nitric oxide, due to its very low
concentration in the residual gas of a spark-ignition engine operated under close to
stoichiometric conditions. Heat losses were also found to strongly influence autoignition
in a physically intuitive manner.

It therefore appears that an accurate model of LPG autoignition in a spark-ignition


engine needs to model several key phenomena. This work demonstrated that such a model
in particular should include careful modelling of the in-cylinder heat transfer and the
residual gas nitric oxide concentration, in addition to the flame propagation modelling and
the autoignition chemistry. When these key phenomena were included in the model, the
autoignition onset for fuels primarily comprised of propane, n-butane and/or iso-butane
could generally be modelled to within 1.0 crank angle degree of experiment. The same
agreement could not be obtained for propylene, and thus further work on the autoignition
chemistry of this hydrocarbon in particular appears to be required.
7.2 Recommendations for future research 191

7.2 Recommendations for future research

The following recommendations are made for future research. The first recommendation
relates to the measured octane numbers of LPGs. The second recommendation relates to
the performance of LPGs in modern, spark-ignition engines. Finally, the third recommen-
dation relates to the detailed autoignition chemistry of LPGs.

1. Extend the octane number data, along with the models that relate fuel composition
to the RON/MON, to include ethane and the butylenes.

This study examined the four constituents that typically represent the major species in all
LPG fuels globally. However, commercial LPGs can also contain several other hydrocar-
bons, which exist in smaller concentrations. The most important of these appears to be the
butylenes (1-butylene, 2-butylene, iso-butylene), with proprietary MSDSs suggesting that
some refinery-sourced LPGs may be comprised of up to 15 (% vol) of these species. The
literature review indicated that some of the butylenes are likely to have MONs that are
below 80.0, and therefore considerably lower than most of the other constituents. Further,
there is some uncertainty surrounding the blending characteristics exhibited by these
species in octane number. It is possible that small concentrations of these species may
cause a notable reduction in the MON of commercial LPGs.

In the case of LPGs produced during the processing of natural gas, the proprietary
MSDSs suggest that some LPGs may also contain up to 10 (% vol) ethane. Although ethane
has a marginally higher RON and MON than propane (American Petroleum Institute,
1958), the limited MON data for ethane/propane mixtures indicates that these constituents
blend antagonistically at small ethane concentrations. The significance of small ethane
concentrations on the octane rating of LPGs should therefore also be established.

2. Evaluate the performance of a range of LPG blends in a modern, spark-ignition


engine, that delivers liquid fuel into the port or directly into the cylinder.

As identified in the literature review, several studies have shown that the RON appears to
192 Conclusions and recommendations for future research

be most relevant to modern, spark-ignition engines. Further, these studies suggest that
fuels of a given RON tend to exhibit the best anti-knock performance with decreasing
MON (i.e. fuels with high octane sensitivities).

If this is the case for LPGs, then there are several mixtures that may offer favourable
anti-knock performance in modern engines, but which do not satisfy the existing fuel
quality standards, due to their low MON. Once such example is a 50/50 (% mol) mixture
comprised of propane and propylene. This fuel has a MON below the 90.5 requirement of
the existing Australian LPG fuel standard, but has a RON that is approximately 12.0 units
higher than a standard gasoline (RON 103.5). If the anti-knock performance of such fuels
in a modern engine was superior to that of a standard gasoline, this would provide strong
evidence to suggest that relaxation of the existing LPG fuel standards is possible.

Further, LPG fuels with equivalent RONs, but with different MONs (and therefore dif-
ferent fuel sensitivities) should also be compared in terms of their anti-knock performance.
This information could be used to establish whether the anti-knock performance of LPGs
in modern engines also appears to correlate most closely with the RON, as is the case for
gasolines. This would provide further evidence in support of using the RON in LPG fuel
standards.

3. Extend and validate the detailed chemical kinetic mechanism to include specific
hydrocarbon-NOx reactions for C3 -C4 paraffins and olefins.

The chemical kinetic mechanism used in this thesis did not contain any specific
hydrocarbon-NOx reactions for molecules larger than ethane. The absence of these re-
actions did not appear to adversely affect the agreement between the modelled and
measured autoignition behaviour for fuels predominantly comprised of the paraffins.
However, there was some evidence to indicate that the absence of specific hydrocarbon-
NOx reactions for propylene could be contributing to the poor agreement between the
modelled and measured autoignition behaviour of this fuel.

In addition to sometimes being present in LPG, propylene is an important intermediate


product that is formed during the oxidation of larger hydrocarbons. Thus, an accurate
7.2 Recommendations for future research 193

description of propylene oxidation is also necessary in order to study the autoignition


behaviour of liquid fuels. It is therefore recommended that a range of specific hydrocarbon-
NOx reactions for propylene first be incorporated into the existing kinetic model. The
equivalent reactions for the C3 -C4 paraffins and C4 olefins should then be added for
consistency.
Bibliography
P. Abbasi Atibeh. Performance of a Spark ignition, Lean burn, Natural Gas Internal
Combustion Engine. Master’s thesis, Department of Mechanical Engineering, The
University of Melbourne, 2012.

W.E. Adams and K. Boldt. What Engines Say About Propane Fuel Mixtures. SAE Technical
Paper 640833., 1964.

S. Addagarla. Preignition Chemical Reactions Leading to Autoignition and Knock in Spark-


Ignition Engines. PhD thesis, Drexel University, 1991.

W.S. Affleck, P.E. Bright, and A. Fish. Run-on in gasoline engines: A chemical description
of some effects of fuel composition. Combust. Flame, 12(4):307–317, 1968.

T.A. Albahri. Structural Group Contribution Method for Predicting the Octane Number of
Pure Hydrocarbon Liquids. Ind. Eng. Chem. Res., 42:657–662, 2003.

L.F. Albright and R.E. Eckert. New equations help rapidly determine alkylate octane
numbers. Oil Gas J., pages 51–60, 1999.

A.C. Alkidas. Combustion-chamber crevices: the major source of engine-out hydrocarbon


emissions under fully warmed conditions. Prog. Energy Combust. Sci., 25(3):253–273,
1999.

American Petroleum Institute. Knocking Characteristics of Pure Hydrocarbons; API Research


Project 45, 1958. ASTM International Special Technical Publication No. 225, Philadelphia,
PA.

J.M. Anderlohr, R. Bounaceur, A. Pires Da Cruz, and F. Battin-Leclerc. Modeling of au-


toignition and NO sensitization for the oxidation of IC engine surrogate fuels. Combust.
Flame, 156:505–521, 2009.

195
196 BIBLIOGRAPHY

J. Andrae, D. Johansson, P. Bjornbom, P. Risberg, and G.T. Kalghatgi. Co-oxidation in the


auto-ignition of primary reference fuels and n-heptane/toluene blends. Combust. Flame,
140(4), 2005.

G. Andrews. Environmental impact and health concerns of SI emissions. Short course on


SI engine emissions, University of Leeds, November 2001.

D. Andrianov. Minimising Cold Start Fuel Consumption and Emissions from a Gasoline
Fuelled Engine. PhD thesis, Department of Mechanical Engineering, The University of
Melbourne, 2011.

W.J.D. Annand. Heat transfer in the cylinders of reciprocating internal combustion engines.
Proc. I. Mech. E., 177(36):973–996, 1963.

W.J.D. Annand and T. Ma. Instantaneous Heat Transfer Rates to the Cylinder Head Surface
of a Small Compression Ignition Engine. Proc. I. Mech. E., 185(72):976–987, 1971.

R. Arslan, Y. Ulusoy, Y. Tekin, and A. Surmen. An evaluation of the alternative transport


fuel policies for Turkey. Energ. Policy, 38(6), 2010.

ASTM International. Standard Test Method for Knock Characteristics of Liquefied Petroleum
(LP) Gases by the Motor (LP) Method, 1986. Designation D2623, West Conshohocken, PA.

ASTM International. Standard Practice for Calculation of Certain Physical Properties of Lique-
fied Petroleum (LP) Gases from Compositional Analysis, 2007a. Designation D2598, West
Conshohocken, PA.

ASTM International. Standard Practice for Interconversion of Analysis of C5 and Lighter


Hydrocarbons to Gas-Volume, Liquid-Volume, or Mass Basis, 2007b. Designation D2421,
West Conshohocken, PA.

ASTM International. Standard Specification for Liquefied Petroleum (LP) Gases, 2011. Designa-
tion D1835, West Conshohocken, PA.

ASTM International. Standard Test Method for Research Octane Number of Spark-Ignition
Engine Fuel, 2012a. Designation D2699, West Conshohocken, PA.
BIBLIOGRAPHY 197

ASTM International. Standard Test Method for Motor Octane Number of Spark-Ignition Engine
Fuel, 2012b. Designation D2700, West Conshohocken, PA.

A.A Attar and G.A. Karim. Knock Rating of Gaseous Fuels. J. Eng. Gas Turbines Power,
125:500–504, 2003.

W.P. Attard. Small Engine Performance Limits- Turbocharging, Combustion and Design. PhD
thesis, Department of Mechanical Engineering, The University of Melbourne, 2007.

C.J. Aul, W.K. Metcalfe, S.M. Burke, H.J. Curran, and E.L. Petersen. Ignition and kinetic
modeling of methane and ethane fuel blends with oxygen: A design of experiments
approach. Combust. Flame, 160(7):1153–1167, 2013.

AusIndustry. LPG Vehicle Scheme Statistics, February 2013. URL http:


//www.ausindustry.gov.au/programs/energy-fuels/lpgvs/Pages/
default.aspx. Canberra, Australia. Accessed 05/04/2013.

Australian Government, Bureau of Resources and Energy Economics. Australian


Petroleum Statistics. Technical report, June 2013a. Canberra, Australia.

Australian Government, Bureau of Resources and Energy Economics. Energy in Australia.


Technical report, May 2013b. Canberra, Australia.

Australian Government, Department of Environment and Heritage. Setting Na-


tional Fuel Quality Standards: Proposed Standards for Liquefied Petroleum Gas (Au-
togas), October 2001. URL environment.gov.au/atmosphere/fuelquality/
publications/\newlinepubs/paper5.pdf. Canberra, Australia. Accessed
01/12/2012.

Australian Liquefied Petroleum Gas Association. Liquefied Petroleum Gas for Automotive
Use Specification, January 2000. Sydney, Australia.

C. Bahrini, O. Herbinet, P. Glaude, C. Schoemaecher, C. Fittschen, and F. Battin-Leclerc.


Quantification of Hydrogen Peroxide during the Low-Temperature Oxidation of Alka-
nes. J. Am. Chem. Soc., 134(29):11944–11947, 2012.

P. Baker and H.C. Watson. MPI air/fuel mixing for gaseous and liquid LPG. SAE Technical
Paper 2005-01-0246., 2005.
198 BIBLIOGRAPHY

J.S. Balis. Personal communication. GE Energy Waukesha, Personal Communication


(13/06/2012)., 2012.

J.S. Balis. Personal communication. GE Energy Waukesha, Personal Communication


(18/07/2013)., 2013.

M. Baratta and E. Spessa. Numerical Simulation Techniques for the Prediction of Fluid-Dynamics,
Combustion and Performance in IC Engines Fuelled by CNG. In: Computational Simulations
and Applications, pages 259–286. InTech, 2011. ISBN 978-953-307-430-6.

M. Baratta, S. d’Ambrosio, E. Spessa, and A. Vassallo. Analysis of Cyclic Variability in


a Bi-Fuel Engine by Means of a Cycle-Resolved Diagnostic Technique. ASME Internal
Combustion Engine Division Fall Technical Conference, pages 175–191, 2005. Paper No.
ICEF2005-1214.

V.Y. Basevich, A.A. Belyaev, and S.M. Frolov. The Mechanisms of Oxidation and Com-
bustion of Normal Alkane Hydrocarbons: The Transition from C1 -C3 to C4 H10 . Russian
Journal of Physical Chemistry B, Focus on Physics, 1(5):477–484, 2007.

J.K. Bechtold and M. Matalon. The dependence of the Markstein length on stoichiometry.
Combust. Flame, 127(1-2):1906–1913, 2001.

N.G. Becker. Models for the response of a mixture. J. Roy. Statist. Soc. Ser. B, 30 (2):349–358,
1968.

D.J. Beerer and V.G. McDonell. An experimental and kinetic study of alkane autoignition
at high pressures and intermediate temperatures. Proc. Combust. Inst, 33(1):301–307,
2011.

A.G. Bell. The Relationship Between Octane Quality and Octane Requirement. SAE
Technical Paper 750935., 1975.

A.B. Bendtsen, P. Glarborg, and K. Dam-Johansen. Low temperature oxidation of methane:


the influence of nitrogen oxides. Combust. Sci. Tech., 151:31–71, 2000.

J.D. Benson and R.F. Stebar. Effects of Charge Dilution on Nitric Oxide Emission from a
Single-Cylinder Engine. SAE Technical Paper 710008, 1971.
BIBLIOGRAPHY 199

A. Bhave, M. Kraft, F. Mauss, A. Oaklet, and H. Zhao. Evaluating the EGR-AFR Operating
Range of a HCCI Engine. SAE Technical Paper 2005-01-0161, 2005.

N. Blin-Simiand, R. Rigny, V. Viossat, S. Circan, and K. Sahetchian. Autoignition of Hydro-


carbon/Air Mixtures in a CFR Engine: Experimental and Modeling Study. Combust. Sci.
Tech., 88(5-6):329–348, 1993.

K. Boldt. Motor (LP) knock test method development. SAE Technical Paper 670055., 1967.

J. Bood, P. Bengtsson, F. Mauss, K. Burgdorf, and I. Denbratt. Knock in Spark-Ignition


Engines: End-Gas Temperature Measurements Using Rotational CARS and Detailed
Kinetic Calculations of the Autoignition Process. SAE Technical Paper 971669, 1997.

G. Borman and K. Nishiwaki. Internal-combustion engine heat transfer. Prog. Energ.


Combust., 13(1):1–46, 1987.

M. Bos. Validation GT-Power Model Cyclops Heavy Duty Diesel Engine. Master’s thesis,
Technische Universiteit Eindhoven, August 2007.

Bosch GmbH. Lambda Sensor LSU 4.9 technical data sheet, April 2011. URL
www.bosch-motorsport.de/pdf/sensors/lambda/lsu49.pdf. Accessed
19/02/2013.

R.A. Bottenberg and J.H. Ward. Applied multiple linear regression, 1963. Technical Report
PRL-TDR-63-6, Lackland AFB, San Antonio, TX.

C.T. Bowman. Kinetics of pollutant formation and destruction in combustion. Prog. Energ.
Combust., 1(1):33–45, 1975.

T.A. Boyd. 1937 Road Knock Tests, Report from the Cooperative Fuel Research Committee.
SAE Trans., 42 (6):244–252, June 1938.

D. Bradley and C. Morley. Autoignition in spark-ignition engines. In: Comprehensive Chemical


Kinetics, volume 35, pages 661–760. Elsevier Science B.V., Amsterdam, The Netherlands,
1997. ISBN 978-0-444-82485-1.
200 BIBLIOGRAPHY

D. Bradley and J. Yeo. A Five-Equation, Four Species Scheme for Modelling Autoignition
of Hydrocarbons. Proceedings of the Anglo-German Combustion Symposium, pages 356–359,
1993.

D. Bradley, G.T. Kalghatgi, C. Morley, P. Snowdon, and J. Yeo. CARS Temperature


Measurements and the Cyclic Dispersion of Knock in Spark Ignition Engines. Proc.
Combust. Inst, 25:125–133, 1994.

R. Braziel. Impact of shale liquids on regional propane supply. Presented at NPGA Winter
Board of Directors Meeting, Key West, FL. Accessed 09/12/2012, January 2012. URL
npga.org/files/public/.

British Petroleum PLC. BP Energy Outlook 2030. Technical report, January 2013. London,
UK.

J.H. Bromly. The Formation of Nitrogen Oxides in Gas Combustion. PhD thesis, Department
of Engineering, Murdoch University, 1991. Perth, Western Australia.

J.H. Bromly, F.J. Barnes, R. Mandyczewsky, T.J. Edwards, and B.S. Haynes. An experimen-
tal investigation of the mutually sensitised oxidation of nitric oxide and n-butane. Proc.
Combust. Inst, 24(1):899–907, 1992.

M. Brower, E.L. Petersen, W. Metcalfe, H.J. Curran, M. Furi, G. Bourque, N. Aluri, and
F. Guthe. Ignition Delay Time and Laminar Flame Speed Calculations for Natural
Gas/Hydrogen Blends at Elevated Pressures. J. Eng. Gas Turbines Power, 135(2):021504,
2013.

S. Brussovansky, J.B. Heywood, and J.C. Keck. Predicting the Effects of Air and Coolant
Temperature, Deposits, Spark Timing and Speed on Knock in Spark Ignition Engines.
SAE Technical Paper 922324, 1992.

S. Burke and H.J. Curran. Uncertainties in rate constants of important reactions for propene
oxidation. 4th International Workshop on Model Reduction in Reacting Flows (IWMRRF),
2013. June 19-21, San Francisco, CA.

A.A. Burluka, K. Liu, C.G.W. Sheppard, A.J. Smallbone, and R. Woolley. The Influence of
BIBLIOGRAPHY 201

Simulated Residual and NO Concentrations on Knock Onset for PRFs and Gasolines.
SAE Technical Paper 2004-01-2998, 2004.

A. Cairns and H. Blaxill. The Effects of Combined Internal and External Exhaust Gas
Recirculation on Gasoline Controlled Auto-Ignition. SAE Technical Paper 2005-01-0133,
2005.

M. Campbell, L. Wyszyński, and R Stone. Combustion of LPG in a Spark-Ignition Engine.


SAE Technical Paper 2004-01-0974., 2004.

A.E. Catania, D. Misul, A. Mittica, and E. Spessa. Unsteady Convection Model for Heat
Release Analysis of IC Engine Pressure Data. SAE Technical Paper 2000-01-1265, 2000.

A.E. Catania, D. Misul, A. Mittica, and E. Spessa. A Refined Two-Zone Heat Release
Model for Combustion Analysis in SI Engines. JSME Int. J. Ser. B, 46(1):75–85, 2003.

J.A. Caton and J.B. Heywood. An Experimental and Analytical Study of Heat Transfer in
an Engine Exhaust Port. Int. J. Heat Mass Transfer, 24(4):581–595, 1981.

N. Cavina, C. Siviero, and R. Suglia. Residual Gas Fraction Estimation: Application to


a GDI Engine with Variable Valve Timing and EGR. SAE Technical Paper 2004-01-2943,
2004.

K. Chakravarthy, R. Wagner, and S. Daw. On the Use of Thermodynamic Modeling


for Predicting Cycle-to-Cycle Variations in a SI Engine under Lean Conditions. SAE
Technical Paper 2005-01-3802, 2005.

Y.L. Chan, F.J. Barnes, J.H. Bromly, A.A. Konnov, and D.K. Zhang. The differentiated effect
of NO and NO2 in promoting methane oxidation. Proc. Combust. Inst, 33:441–447, 2011.

Y.L. Chan, J.H. Bromly, A.A. Konnov, and D.K. Zhang. The Comparative and Combined
Effects of Nitric Oxide and Higher Alkanes in Sensitizing Methane Oxidation. Combust.
Sci. Tech., 184:114–132, 2012.

J. Chang, O. Guralp, Z. Filipi, D. Assanis, T.-W. Kuo, P. Najt, and R. Rask. New Heat
Transfer Correlation for an HCCI Engine Derived from Measurements of Instantaneous
Surface Heat Flux. SAE Technical Paper 2004-01-2996, 2004.
202 BIBLIOGRAPHY

R. Chiriac, B. Radu, N. Apostolescu, and D. Fuiorescu. An Investigation of Knock in a


Spark Ignition Engine Using LPG. SAE Technical Paper 2005-24-027, 2005.

R. Chiriac, B. Radu, and N. Apostolescu. Defining Knock Characteristics and Autoignition


Conditions of LPG with a Possible Correlation for the Control Strategy in a SI Engine.
SAE Technical Paper 2006-01-0227, 2006.

H. Cho, K. Lee, J. Lee, J. Yoo, and K. Min. Measurements and Modeling of Residual Gas
Fraction in SI Engines. SAE Technical Paper 2001-01-1910, 2001.

S.L. Chow. Statistical Significance: Rationale, Validity and Utility. Sage, London, UK, first
edition, 1997. ISBN 9780761952053.

K.M. Chun, J.B. Heywood, and J.C. Keck. Prediction of Knock Occurrence in a Spark-
Ignition Engine. Proc. Combust. Inst, 22:455–463, 1988.

E.A. Clifford. Practical Guide to LP-Gas Utilization. Harbrace Publications, Duluth, MN,
fourth edition, 1969.

R.L.J. Coetzer, R. Rossouw, A. Swarts, and C. Viljoen. The estimation of knock-points of


fuels by a weighted mean square error criterion. Fuel, 85(12-13):1880–1893, 2006.

G. Colin, P. Giansetti, Y. Chamaillard, and P. Higelin. In-Cylinder Mass Estimation using


Cylinder Pressure. SAE Technical Paper 2007-24-0049, 2007.

Commonwealth Scientific and Industrial Research Organisation. Fuel for thought. The
future of transport fuels: challenges and opportunities. Technical report, June 2008.
Canberra, Australia.

F. Contino, F. Foucher, P. Dagaut, T. Lucchini, G. D’Errico, and C. Mounaim-Rousselle.


Experimental and numerical analysis of nitric oxide effect on the ignition of iso-octane
in a single cylinder HCCI engine. Combust. Flame, 160(8):1476–1483, 2013.

J.A. Cornell. Experiments with Mixtures. Designs, Models, and the Analysis of Mixture Data.
Wiley, Hoboken, NJ, third edition, 2002.
BIBLIOGRAPHY 203

H.J. Curran, P. Gaffuri, W.J. Pitz, C.K. Westbrook, and W.R. Leppard. Autoignition
Chemistry of the Hexane Isomers: An Experimental and Kinetic Modeling Study. SAE
Technical Paper 952406, 1995.

H.J. Curran, P. Gaffuri, W.J. Pitz, and C.K. Westbrook. A Comprehensive Modeling Study
of n-Heptane Oxidation. Combust. Flame, 114:149–177, 1998.

H.J. Curran, P. Gaffuri, W.J. Pitz, and C.K. Westbrook. A Comprehensive Modeling Study
of iso-Octane Oxidation. Combust. Flame, 129:253–280, 2002.

P. Dagaut and G. Dayma. The high-pressure reduction of nitric oxide by a natural gas
blend. Combust. Flame, 143(1-2), 2005.

P. Dagaut and K. Hadj Ali. Kinetics of oxidation of a LPG blend mixture in a JSR: experi-
mental and modeling study. Fuel, 82:475–480, 2003.

P. Dagaut and A. Nicolle. Experimental study and detailed kinetic modeling of the effect
of exhaust gas on fuel combustion: mutual sensitization of the oxidation of nitric oxide
and methane over extended temperature and pressure ranges. Combust. Flame, 140:
161–171, 2005.

P. Dagaut, J. Luche, and M. Cathonnet. Reduction of NO by propane in a JSR at 1 atm:


experimental and kinetic modeling. Fuel, 80:979–986, 2001.

B.A. D’Alleva and W.G. Lovell. Relation of Exhaust Gas Composition to Air-Fuel Ratio.
SAE Technical Paper 360106, 1936.

W. A. Daniel. Engine Variable Effects on Exhaust Hydrocarbon Composition (A Single-


Cylinder Engine Study With Propane as the Fuel). SAE Technical Paper 670124, 1967.

R. Danielis and A. Chiabai. Estimating the cost of air pollution from road transport in
Italy. Transport. Res. Part D, 3(4):249–258, 1998.

L. Danis. Engine Valve Cooling. SAE Technical Paper 730055, 1973.

R.S. Davis and G.J. Patterson. Cylinder Pressure Data Quality Checks and Procedures to
Maximize Data Accuracy. SAE Technical Paper 2006-01-1346, 2006.
204 BIBLIOGRAPHY

S.G. Davis and C.K. Law. Determination of and Fuel Structure Effects on Laminar Flame
Speeds of C1 to C8 Hydrocarbons. Combust. Sci. Tech., 140(1-6):427–449, 1998.

J. Dec, M. Sjoberg, and W. Hwang. The Effects of EGR and Its Constituents on the
Autoignition of Single- and Two-Stage Fuels. Presented at 13th Diesel Engine-Efficiency
and Emissions Research Conference, Detroit, MI., 13-16 August 2007.

J. Demuynck, N. Raes, M. Zuliani, M. De Paepe, R. Sierens, and S. Verhelst. Local heat


flux measurements in a hydrogen and methane spark ignition engine with a thermopile
sensor. Int. J. Hydrogen Energ., 34:9857–9868, 2009.

P. Dennis and B. van Deventer. Design and Development of a Hydrogen-Fuelled Internal


Combustion Engine. Technical report, Department of Mechanical Engineering, The
University of Melbourne, October 2008.

G. Deslandes. The motored engine autoignition of hydrogen and methane. PhD thesis, Depart-
ment of Mechanical and Manufacturing Engineering, The University of Melbourne.,
1975.

M. Diesendorf, D. Lamb, J. Mathews, and G. Pearman. A Roadmap for Alternative Fuels


in Australia: Ending our Dependence on Oil. Technical report, July 2008. Report of
Jamison Group to NRMA Motoring and Services. Sydney, Australia.

S.D. Dimitrov and D.I. Kamenski. Binary mixture properties: fitting with canonical rational
functions. Comput. Chem. Eng., 23 (8):1011–1019, 1999.

P. Dimpelfeld and D. Foster. The prediction of auto ignition in a spark-ignited engine.


SAE Technical Paper 841337, 1984.

G.G. Dober. Geometric Control of Hydrocarbon Emissions. PhD thesis, Department of


Mechanical and Manufacturing Engineering, The University of Melbourne, 2002.

A. Douaud and P. Eyzat. DIGITAP-An On-Line Acquisition and Processing System for
Instantaneous Engine Data-Applications. SAE Technical Paper 770218, 1977.

D. Downs, A.D. Walsh, and R.W. Wheeler. A Study of the Reactions that Lead to ‘Knock’
in the Spark-Ignition Engine. Phil. Trans. R. Soc. Lond. A, 243:463–524, 1951.
BIBLIOGRAPHY 205

N.R. Draper and F. Pukelsheim. Mixture models based on homogeneous polynomials. J.


Statist. Plann. Inference, 71 (1-2):303–311, 1998.

A. Dreizler and U. Mass. Influence of the Temperature Distribution on the Auto-Ignition


in the End Gas of Otto engines. Fourth Int. Symp. COMODIA, pages 197–202, 1998.

A. Dubreuil, F. Foucher, and C. Mounaim-Rousselle. Effect of EGR Chemical Components


and Intake Temperature on HCCI Combustion Development. SAE Technical Paper
2006-32-0044, 2006.

A. Dubreuil, F. Foucher, C. Mounaim-Rousselle, G. Dayma, and P. Dagaut. HCCI combus-


tion: Effect of NO in EGR. Proc. Combust. Inst, 31:2879–2886, 2007.

J. Dutczak and K. Golec. Spark Ignition Engine Fuelled by Means of Liquid Propane-
Butane Injection. SAE Technical Paper 2002-01-1689, 2002.

W.L. Easley, A. Agarwal, and G.A. Lavoie. Modeling of HCCI Combustion and Emissions
Using Detailed Chemistry. SAE Technical Paper 2001-01-1029, 2001.

P. Eckert, S. Kong, and R.D. Reitz. Modeling Autoignition and Engine Knock Under Spark
Ignition Conditions. SAE Technical Paper 2003-01-0011, 2003.

E. Edgar. Measurement of Knock Characteristics of Gasoline in Terms of a Standard Fuel.


Ind. Eng. Chem., 19(1):145–146, January 1927.

G. Eichelberg. Some New Investigations on Old Combustion-Engine Problems. Engineer-


ing, 149, 1939.

R.J. Ellison, G.A. Harrow, and B.M. Hayward. The Effect of Tetraethyl-Lead on Flame
Propagation and Cyclic Dispersion in Spark-Ignition Engines. Journal of the Institute of
Petroleum, 54(537):243–250, September 1968.

J.A. Eng, W.R. Leppard, and T.M. Sloane. The Effect of POx on the Autoignition Chemistry
of n-Heptane and Isooctane in an HCCI Engine. SAE Technical Paper 2002-01-2861, 2002.

European Committee for Standardisation. Automotive fuels- LPG- Requirements and test
methods, 2008. Designation EN 589, Brussels, Belgium.
206 BIBLIOGRAPHY

R.J. Falkiner. Liquefied Petroleum Gas. In: Fuels and Lubricants Handbook, pages 31–59. ASTM
International, West Conshohocken, PA, 2003. ISBN 9780803120969.

T. Faravelli, P. Gaffuri, E. Ranzi, and J.F. Griffiths. Detailed thermokinetic modelling of


alkane autoignition as a tool for the optimization of performance of internal combustion
engines. Fuel, 77(3):147–155, 1998.

Federal Register of Legislative Instruments. Fuel Standard (Petrol) Determination,


2001. URL http://www.environment.gov.au/atmosphere/fuelquality/
standards/petrol.html. Canberra, Australia. Accessed 01/12/2012.

Federal Register of Legislative Instruments. Fuel Standard (Autogas) Determination,


2003. URL http://www.environment.gov.au/atmosphere/fuelquality/
standards/lpg.html. Canberra, Australia. Accessed 01/12/2012.

I. Finlay, D. Boam, J. Bingham, and T. Clark. Fast Response FID Measurement of Unburned
Hydrocarbons in the Exhaust Port of a Firing Gasoline Engine. SAE Technical Paper
902165, 1990.

J. Fitton and R. Nates. Knock Erosion in Spark-Ignition Engines. SAE Technical Paper
962102, 1996.

S.B. Fiveland and D.N. Assanis. Development of a Two-Zone HCCI Combustion Model
Accounting for Boundary Layer Effects. SAE Technical Paper 2001-01-1028, 2001.

D. Flowers, S. Aceves, R. Smith, J. Torres, J. Girard, and R. Dibble. HCCI in a CFR Engine:
Experiments and Detailed Kinetic Modeling. SAE Technical Paper 2000-01-0328, 2000.

T. Foong, K.J. Morganti, M.J. Brear, G. da Silva, Y. Yang, and F.L. Dryer. The Effect of
Charge Cooling on the RON of Ethanol/Gasoline Blends. SAE Int. J. Fuels Lubr., 6(1):
34–43, 2013.

S.B. Gallas. Characteristics Of An ASTM-CFR Engine Operating With A Propane Fuel


System. Master’s thesis, Department of Mechanical Engineering, University of Missouri,
May 1959.
BIBLIOGRAPHY 207

F. Galliot, W.K. Cheng, C. Cheng, M. Sztenderowicz, J.B. Heywood, and N. Collings.


In-Cylinder Measurements of Residual Gas Concentration in a Spark Ignition Engine.
SAE Technical Paper 900485, 1990.

E. Galloni, G. Fontana, and R. Palmaccio. Numerical analyses of EGR techniques in a


turbocharged spark-ignition engine. Appl. Therm. Eng., 39:95–104, 2012.

Gamma Technologies Inc. GT-Power User’s Manual, version 7.3, 2012a. Westmont, IL.

Gamma Technologies Inc. GT-Suite Flow Theory Manual, version 7.3, 2012b.

J.H. Gary, G.E. Handwerk, and M.J. Kaiser. Petroleum Refining. CRC Press, Boca Raton, FL,
fifth edition, 2007. ISBN 0-8493-7038-8.

Gas Processors Association. Liquefied Petroleum Gas Specifications and Test Methods, 1997.
Designation 2140, Tulsa, OK.

S. Gersen, H. Darmeveil, and H. Levinsjy. The effects of CO addition on the autoignition


of H2, CH4 and CH4/H2 fuels at high pressure in an RCM. Combust. Flame, 159(12):
3472–3475, 2012.

P. Ghosh, K.J. Hickey, and S.B. Jaffe. Development of a detailed gasoline composition-based
octane model. Ind. Eng. Chem. Res., 45:337–345, 2006.

R. Gist. Developments in international supply and demand of propane. Presented at


NPGA Winter Board of Directors Meeting, Key West, FL. Accessed 09/12/2012, January
2012. URL npga.org/files/public/.

M. Gluckstein and C. Walcutt. End-Gas Temperature-Pressure Histories and Their Relation


To Knock. SAE Technical Paper 610044, 1961.

C.J. Goy, A.J. Moran, and G.O. Thomas. Autoignition Characteristics of Gaseous Fuels at
Representative Gas Turbine Conditions. American Society of Mechanical Engineers (ASME)
51st All International Gas Turbine and Aeroengine Congress and Exhibition, 2001.

P. Graham, L. Reedman, and F. Poldy. Modelling of the future of transport fuels. Technical
report, Commonwealth Scientific and Industrial Research Organisation (CSIRO), June
2008. Report IR 1046. Canberra, Australia.
208 BIBLIOGRAPHY

R. Green, N. Cernansky, W. Pitz, and C. Westbrook. The Role of Low Temperature


Chemistry in the Autoignition of n-Butane. SAE Technical Paper 872108, 1987a.

R.M. Green, C.D. Parker, W.J. Pitz, and C.K. Westbrook. The Autoignition of Isobutane in
a Knocking Spark Ignition Engine. SAE Technical Paper 870169, 1987b.

Y.G. Guezennec and W. Hamama. Two-Zone Heat Release Analysis of Combustion Data
and Calibration of Heat Transfer Correlation in an I. C. Engine. SAE Technical Paper
1999-01-0218, 1999.

E. Hairer and G. Wanner. Solver RADAU5, January 2002. URL http://www.unige.ch/


˜hairer/prog/stiff/radau.f.

S. Hajireza, F. Mauss, and B. Sunden. Two-Zone Model of Gas Thermodynamic State in SI


Engines with Relevance foe Knock. Fourth Int. Symp. COMODIA, pages 203–208, 1998.

S. Hajireza, B. Sunden, and F. Mauss. A Three-Zone Model for Investigation of Gas


Behavior in the Combustion Chamber of SI Engines in Relation to Knock. SAE Technical
Paper 1999-01-0219, 1999.

K.I. Hamada, M.M. Noor, K. Kadirgama, and R.A. Bakar. Effect of intake conditions
on heat transfer characteristics for port injection hydrogen fueled engine. 2nd Interna-
tional Conference on Mechanical and Electrical Technology (ICMET), pages 549–554, 10-12
September 2010.

T.K. Hanson and A.C. Egerton. Nitrogen Oxides in Internal Combustion Engine Gases.
Proc. Roy Soc. A., 163(912):90–100, 1937.

W.M. Hart. Impact of shale gas development on NGL supplies


and petrochemical feedstocks. Presented at South Texas Section of
AIChE, Houston, TX. Accessed 09/12/2012, February 2012. URL
aiche.org/resources/chemeondemand/webinars/impact-shale-\
newlinegas-development-ngl-supplies-and-petrochemical-feedstocks.

T.K. Hayes, White R.A., and Peters J.E. Combustion Chamber Temperature and Instanta-
neous Local Heat Flux Measurements in a Spark Ignition Engine. SAE Technical Paper
930217, 1993.
BIBLIOGRAPHY 209

D. Healy, N.S. Donato, C.J. Aul, E.L. Petersen, C.M. Zinner, G. Bourque, and H.J. Curran.
nButane ignition delay time measurements at high pressure and detailed chemical
kinetic simulations. Combust. Flame, 157:1526–1539, 2010a.

D. Healy, N.S. Donato, C.J. Aul, E.L. Petersen, C.M. Zinner, G. Bourque, and H.J. Curran.
Isobutane ignition delay time measurements at high pressure and detailed chemical
kinetic simulations. Combust. Flame, 157:1540–1551, 2010b.

D. Healy, D.M. Kalitan, C.J. Aul, E.L. Petersen, G. Bourque, and H.J. Curran. Oxidation of
C1-C5 Alkane Quinternary Natural Gas Mixtures at High Pressures. Energ. Fuel, 24(3):
1521–1528, 2010c.

D. Healy, M.M. Kopp, N.L. Polley, E.L. Petersen, G. Bourque, and H.J. Curran. Methane/n-
Butane Ignition Delay Measurements at High Pressure and Detailed Chemical Kinetic
Simulations. Energ. Fuel, 24(3):1617–1627, 2010d.

S. Heimel and R.C. Weast. Effect of Initial Mixture Temperature on the Burning Velocity of
Benzine-Air, n-Heptane-Air and iso-Octane-Air Mixtures. Proc. Combust. Inst, 6, 1957.

P.R. Hellier. The Molecular Structure of Future Fuels. PhD thesis, University College London,
University of London., March 2013.

F. Hendren. Propane power for light duty vehicles: An overview. SAE Technical Paper
830383., 1983.

J.B. Heywood. Internal Combustion Engine Fundamentals. McGraw-Hill Science Engineering,


NY, USA, 1988. ISBN 0-07-028637-X.

J.B. Heywood, J.C. Keck, and J.M. Rife. Hydrocarbon Formation and Oxidation in Spark-
Ignition Engines. Final Report on a Research Program Funded by General Motors Research
Laboratories September 1976 to August 1979, 1980.

J.C. Hilliard and R.W. Wheeler. Nitrogen Dioxide in Engine Exhaust. SAE Technical Paper
790691, 1979.

M. Hori, N. Matsunaga, N. Marinov, W. Pitz, and C. Westbrook. An Experimental and


Kinetic Calculation of the Promotion Effect of Hydrocarbons on the NO-NO2 Conversion
in a Flow Reactor. Proc. Combust. Inst, 27:389–396, 1998.
210 BIBLIOGRAPHY

E. Hu, Z. Huang, X. Jiang, and J. Zhang. Study on Ignition Delay Times of DME and
n-Butane Blends. SAE Technical Paper 2013-01-1146, 2013.

J. Hvezda. Multi-Zone Models of Combustion and Heat Transfer Processes in SI Engines.


SAE Technical Paper 2011-37-0024, 2011.

H.S. Imming. The Effect of Piston-Head Temperature on Knock-Limited Power. Technical


report, NACA-WR-E-35, July 1944. Glenn Research Center, Cleveland, OH.

J. Ingamells and E. Jones. Developing road octane correlations from octane requirement
surveys. SAE Technical Paper 810492., 1981.

Intergovernmental Panel on Climate Change. Contribution of Working Group I to the


Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Technical
report, May 2007. Cambridge, United Kingdom.

International Energy Agency. World Energy Outlook 2012 . Technical report, November
2012. Paris, France.

International Organisation for Standardisation. Commercial propane and butane - Analysis by


gas chromatography, 1988. Designation ISO 7941, Geneva, Switzerland.

International Organisation for Standardisation. Liquefied petroleum gases- Calculation method


for density and vapour pressure, 1997. Designation ISO 8973, Geneva, Switzerland.

A. Iob, M.A. Ali, B.S. Tawabini, J.A. Anabtawi, S.A. Ali, and A. Al-Farayedhi. Prediction
of reformate research octane number by FT-i.r. spectroscopy. Fuel, 74(2):227–231, 1995.

A. Iqbal. Fundamentals of Knock. PhD thesis, The Ohio State University, 2012.

A. Irimescu. Convective heat transfer equation for turbulent flow in tubes applied to
internal combustion engines operated under motored conditions. Appl. Therm. Eng.,
50(1):536–545, 2013.

R.J. Jenkin, E.H. James, and W. Malalasekera. Predicting the Onset of End-Gas Autoignition
with a Quasi-Dimensional Spark Ignition Engine Model. SAE Technical Paper 972877,
1997.
BIBLIOGRAPHY 211

J.H. Johnson, P.S. Myers, and O.A. Uyehara. End-Gas Temperatures, Pressures, Reaction
Rates, and Knock. SAE Technical Paper 650505, 1965.

E.W. Kaiser, W.G. Rothschild, and G.A. Lavoie. The Effect of Fuel and Operating Variables
on Hydrocarbon Species Distributions in the Exhaust from a Multicylinder Engine.
Combust. Sci. and Tech., 32:245–265, 1983.

E.W. Kaiser, W.G. Rothschild, and G.A. Lavoie. Storage and Partial Oxidation of Unburned
Hydrocarbons in Spark-Ignited Engines - Effect of Compression Ratio and Spark Timing.
Combust. Sci. and Tech., 36(3-4):171–189, 1984.

G.T. Kalghatgi. Fuel Anti-Knock Quality- Part I. Engine Studies. SAE Technical Paper
2001-01-3584., 2001a.

G.T. Kalghatgi. Fuel Anti-Knock Quality- Part II. Vehicle Studies- How Relevant is Motor
Octane Number (MON) in Modern Engines? SAE Technical Paper 2001-01-3585., 2001b.

G.T. Kalghatgi, M. Golombok, and P. Snowdon. Fuel Effects on Knock, Heat Release and
CARS Temperatures in a Spark Ignition Engine. Combust. Sci. Tech., 110-111(1):209–228,
1995a.

G.T. Kalghatgi, P. Snowdon, and C. McDonald. Studies of Knock in a Spark Ignition Engine
with CARS Temperature Measurements and Using Different Fuels. SAE Technical Paper
950690, 1995b.

G.T. Kalghatgi, L. Hildingsson, A.J. Harrison, and B. Johansson. Surrogate fuels for
premixed combustion in compression ignition engines. Int. J. Engine Res., 12:1–14, 2011.

D. Kayes and S. Hochgreb. Development of a Time and Space Resolved Sampling Probe
Diagnostic for Engine Exhaust Hydrocarbons. SAE Technical Paper 961002, 1996.

R. Kerley and K. Thurston. Knocking Behaviour of Fuels and Engines. SAE Technical Paper
560049., 1956.

L.J. Kirsch and D.B. Pye. Modelling the Action of Antiknock Agents. Proc. Combust. Inst, 2:
806, 1975.
212 BIBLIOGRAPHY

R.F. Klausmeier and I.F. Billick. Comparative Analysis of the Environmental Impact of
Alternative Transportation Fuels. Energ. Fuel, 7:27–32, 1993.

V. Knop, C. Pera, and F. Duffour. Validation of a ternary gasoline surrogate in a CAI


engine. Combust. Flame, 160(10):2067–2082, 2013.

G. Konig and C. Sheppard. End Gas Autoignition and Knock in a Spark Ignition Engine.
SAE Technical Paper 902135, 1990.

A.A. Konnov. Remaining uncertainties in the kinetic mechanism of hydrogen combustion.


Combust. Flame, 152:507–528, 2008.

A.A. Konnov, F.J. Barnes, J.H. Bromly, J. Zhu, and D. Zhang. The pseudo-catalytic pro-
motion of nitric oxide oxidation by ethane at low temperatures. Combust. Flame, 141:
191–199, 2005.

S. Kowalski, T.J. Held, Y.S. Stein, and F.L. Dryer. Reactivity of an 87 Octane Number Blend
of the Primary Reference Fuels with Added Nitric Oxide. Presented at the Fall Meeting
of the Western States Section/The Combustion Institute, Berkeley, CA, 12-13 October, 1992.
Paper No. WSSSCI 92-91.

J. Kubesh, S.R. King, and W.E. Liss. Effect of Gas Composition on Octane Number of
Natural Gas Fuels. SAE Technical Paper 922359, 1992.

S. Kumagai and Y. Kudo. Flame studies by means of ionization gap in a high-speed


spark-ignition engine. Proc. Combust. Inst, 9(1):1083–1087, 1963.

N.M.M. Lambert. An Experimental Study of Exhaust Hydrocarbon Emissions from a


Spark Ignition Engine. Master’s thesis, Loughborough University of Technology, May
1984.

N. Lamia, L. Wolff, P. Leflaive, P. Sa Gomes, C.A. Grande, and A.E. Rodrigues.


Propane/propylene separation by simulated moving bed I. Adsorption of propane,
propylene and isobutane in pellets of 13X Zeolite. Sep. Sci. Technol., 42 (12):2539–2566,
2007.

D.R. Lancaster, R.B. Krieger, and J.H. Lienesch. Measurement and Analysis of Engine
Pressure Data. SAE Technical Paper 750026, 1975.
BIBLIOGRAPHY 213

D.R. Lancaster, R.B. Krieger, S.C. Sorensen, and W.L. Hull. Effects of Turbulence on
Spark-Ignition Engine Combustion. SAE Technical Paper 760160, 1976.

O. Lang, K. Habermann, R. Thiele, and F. Fricke. Gasoline Combustion with Future Fuels.
SAE Technical Paper 2007-26-021, 2007.

R. Lanzafame. Water Injection Effects In A Single-Cylinder CFR Engine. SAE Technical


Paper 1999-01-0568, 1999.

G.A. Lavoie, J.A. Lorusso, and A.A. Adamnyk. Hydrocarbon Emissions Modeling for Spark
Ignition Engines. In: Combustion Modeling in Reciprocating Engines, pages 409–445. Plenum
Press, 1978.

M. Leiker, K. Christoph, M. Rankl, W. Cartellion, and U. Pfefer. Evaluation of Anti-


Knock Property of Gaseous Fuels by Means of the Methane Number and Its Practical
Application to Gas Engines. ASME Paper 72-DGP-4., pages 1–5, 1972.

W.R. Leppard. A Detailed Chemical Kinetics Simulation of Engine Knock. Combust. Sci.
Tech., 43 (1-2):1–20, 1985.

W.R. Leppard. The Chemical Origin of Fuel Octane Sensitivity. SAE Technical Paper 902137.,
1990.

S.Y. Liao, D.M. Jiang, J. Gao, Z.H. Huang, and Q. Cheng. Measurements of Markstein
numbers and laminar burning velocities for Liquefied Petroleum Gas-air mixtures. Fuel,
83:1281–1288, 2004.

S.Y. Liao, D.M. Jiang, Q. Cheng, J. Gao, Z.H. Huang, and Y. Hu. Correlations for laminar
burning velocities of Liquefied Petroleum Gas-air mixtures. Energ. Convers. Manage., 46:
3175–3184, 2005.

M.A. Liberman, M.F. Ivanov, O.E. Peil, D.M. Valiev, and L.E. Eriksson. Numerical Modeling
of the Propagating Flame and Knock Occurrence in Spark-Ignition Engines. Combust.
Sci. and Tech., 177:151–182, 2005.

M.A. Liberman, M.F. Ivanov, D.M. Valiev, and L.E. Eriksson. Hot Spot Formation by
the Propagating Flame and the Infleunce of EGR on Knock Occurrence in SI Engines.
Combust. Sci. and Tech., 178:1613–1647, 2006.
214 BIBLIOGRAPHY

F. Lindstrom, H. Å ngstrom, G. Kalghatgi, and C.E. Moller. An Empirical SI Combustion


Model Using Laminar Burning Velocity Correlations. SAE Technical Paper 2005-01-2106,
2005.

Z. Liu. Chemical Kinetics Modelling Study on Fuel Autoignition in Internal Combustion Engines.
PhD thesis, Loughborough University, July 2010.

Z. Liu and R. Chen. A Zero-Dimensional Combustion Model with Reduced Kinetics for SI
Engine Knock Simulation. Combust. Sci. Tech., 181:828–852, 2009.

W.G. Lovell. Knocking characteristics of hydrocarbons. J. Ind. Eng. Chem., 40:2388–2438,


1948.

W.G. Lovell, J.M. Campbell, and T.A. Boyd. Knocking characteristics of hydrocarbons
determined from compression ratios at which individual compounds begin to knock
under specified conditions. J. Ind. Eng. Chem., 26 (10):1105–1108, 1934.

W. Lowry, J. De Vries, M. Krejci, E. Petersen, Z. Serinyel, W. Metcalfe, H. Curran, and


G. Bourque. Laminar Flame Speed Measurements and Modeling of Pure Alkanes and
Alkane Blends at Elevated Pressures. J. Eng. Gas Turbines Power, 133(9):091501, 2011.

X. Lu, W. Chen, Y. Hou, and Z. Huang. Study on the Ignition, Combustion and Emissions
of HCCI Combustion Engines Fueled With Primary Reference Fuels. SAE Technical Paper
2005-01-0155, 2005.

R.P. Lucht, R.E. Teets, R.M. Green, R.E. Palmer, and C.R. Furguson. Unburnt Gas Tempera-
tures in an Internal Combustion Engine: I CARS Temperature Measurements. Combust.
Sci. Tech., 55:41–61, 1987.

E.J. Lyford-Pike and J.B. Heywood. Thermal boundary layer thickness in the cylinder of a
spark-ignition engine. Int. J. Heat Mass Transfer, 27(10):1873–1878, 1984.

H. Machrafi, S. Cavadias, and P. Guibert. An experimental and numerical investigation


of the influence of external gas recirculation on the HCCI autoignition process in an
engine: Thermal, diluting, and chemical effects. Combust. Flame, 155:476–489, 2008.

R.R. Maly, R. Klein, N. Peters, and G. Konig. Theoretical and Experimental Investigation
of Knock Induced Surface Destruction. SAE Technical Paper 900025, 1990.
BIBLIOGRAPHY 215

X. Man, C. Tang, L. Wei, L. Pan, and Z. Huang. Measurements and kinetic study on
ignition delay times of propane/hydrogen in argon diluted oxygen. Int. J. Hydrogen
Energy, 38(5):2523–2530, 2013.

N.M. Marinov, W.J. Pitz, C.K. Westbrook, M. Hori, and N. Matsunaga. An Experimental
and Kinetic Calculation of the Promotion Effect of Hydrocarbons on the NO-NO2
Conversion in a Flow Reactor. Proc. Combust. Inst., 27:389–396, 1998a.

N.M. Marinov, W.J. Pitz, C.K. Westbrook, A.M. Vincitore, M.J. Castaldi, and S.M. Senkan.
Aromatic and Polycyclic Aromatic Hydrocarbon Formation in a Laminar Premixed
n-Butane Flame. Combust. Flame, 114:192–213, 1998b.

M.A. Marr, J.S. Wallace, S. Chandra, L. Pershin, and J. Mostaghimi. A fast response
thermocouple for internal combustion engine surface temperature measurements. Exp.
Therm. Fluid Sci., 34:183–189, 2010.

J. Mason and H. Hesselberg. Engine Knock as Influenced by Precombustion Reactions.


SAE Technical Paper 540229, 1954.

H. Menrad, G. Decker, and R. Wegener. The Potential of Methanol and LPG as new fuels
for Transportation. Resour. Conserv., 10:125–133, 1983.

J. Meyer. Innovations in Representation and Calibration of Resdiaul Gas Fraction and


Volumetric Efficiency in a Spark Ignited, Internal Combustion Engine. Master’s thesis,
The Ohio State University, 2008.

Minitab Inc. Minitab, version 16.2.2, 2010. State College, PA.

V. Mittal and J.B. Heywood. The relevance of fuel RON and MON to knock onset in
modern SI engines. SAE Technical Paper 2008-01-2414., 2008.

V. Mittal and J.B. Heywood. The shift in relevance of fuel RON and MON to knock onset
in modern SI engines over the last 70 years. SAE Int. J. Engines, 2 (2):1–10, 2009.

N. Mizushima, S. Sato, Y. Ogawa, T. Yamamoto, U. Sawut, B. Takigawa, K. Kawayoko,


and G. Konagai. Combustion Characteristics and Performance Increase of an LPG-SI
Engine with Liquid Fuel Injection System. SAE Technical Paper 2009-01-2785., 2009.
216 BIBLIOGRAPHY

N. Mizushima, T. Yamamoto, J. Kusaka, and S. Susumu. Study on Burning Velocity of


LPG Fuel in a Constant Volume Combustion Chamber and an SI Engine. SAE Technical
Paper 2010-01-0614, 2010.

MKS Instruments. MKS Type 1500 Series Mass-Flo


R
Controller Instruction Manual, Rev. D,
1996. Andover, MA.

D.C Montgomery. Design and Analysis of Experiments. Wiley, Hoboken, NJ, 8 edition, 2012.
ISBN 1-118-14692-1.

N.P.W. Moore and B.N. Roy. Studies of Autoignition and Knock with Propane as an Engine
Fuel. Combust. Flame, 3:421–435, 1959.

G. Moreac. An Experimental and Modelling Study of Chemical Interactions between Residual


Gases and Fresh Gases in Gasoline Homogeneous Charge Compression Ignition Engines (In
French). PhD thesis, University of Orleans, 2003.

G. Moreac, P. Dagaut, J.F. Roesler, and M. Cathonnet. Nitric oxide interactions with
hydrocarbon oxidation in a jet-stirred reactor at 10 atm. Combust. Flame, 145:512–520,
2006.

K.J. Morganti, T. Foong, M.J. Brear, G. da Silva, and F.L. Dryer. A Comparison of Reference
Engine Autoignition with Detailed Kinetic Modelling for a range of Primary Reference
Fuels. Proc. Aust. Combust. Sym., Nov. 29 - Dec 1, 2011.

E. Moses, A. Yarin, and P. Bar-Yoseph. On Knocking Prediction in Spark Ignition Engines.


Combust. Flame, 101:239–291, 1995.

A. Muller. New Method produces accurate octane blending values. Oil Gas J. Annual
Refining Report, page 80, March 1992.

F. Mustovic. Autogas Propulsion Systems for Automobiles. IBC Engineering and Publishing,
Sarajevo, Bosnia and Herzegovina, 2011. ISBN 978-9958-9173-2-5.

C.V. Naik, K.V. Puduppakkam, and E. Meeks. An Improved Core Reaction Mechanism for
Saturated C0 -C4 Fuels. J. Eng. Gas Turbines Power, 134 (2), 2012.
BIBLIOGRAPHY 217

NASA Glenn Research Center. Thermo Build, NASA Glenn thermodynamic database, 2006.
URL http://www.grc.nasa.gov/WWW/CEAWeb/ceaThermoBuild.htm.

B. Natarajan and F.V. Bracco. On multidimensional modeling of auto-ignition in spark-


ignition engines. Combust. Flame, 57(2):179–197, 1984.

R.J. Nates. Thermal Stresses Induced by Knocking Combustion in Spark-Ignition Engines.


SAE Technical Paper 2000-01-1238, 2000.

R.J. Nates and A.D.B. Yates. Knock Damage Mechanisms in Spark-Ignition Engines. SAE
Technical Paper 942064, 1994.

G.J. Nebel and M.W. Jackson. Some Factors Affecting the Concentration of Oxides of
Nitrogen in Exhaust Gases from Spark Ignition Engines. JAPCA J. Air Waste Ma., 8(3):
213–219, 1958.

W.P. Nel and C.J. Cooper. Implications of fossil fuel constraints on economic growth and
global warming. Energ. Policy, 37:166–180, 2009.

P.F. Nelson and B.S. Haynes. Hydrocarbon-NOX Interactions at Low Temperatures- 1.


Conversion of NO to NO2 Promoted by Propane and the Formation of HNCO. Proc.
Combust. Inst, 25:1003–1010, 1994.

J. Ninomiya and A. Golovoy. Effects of Air-Fuel Ratio on Composition of Hydrocarbon


Exhaust from Isooctane, Diisobutylene, Toluene, and Toluene-n-Heptane Mixture. SAE
Technical Paper 690504, 1969.

K. Nishiwaki. Modeling Engine Heat Transfer and Flame-Wall Interaction. 4th Int. Symp.
COMODIA, pages 35–44, 1998.

P. Obwald, K. Kohse-Hoinghaus, U. Struckmeier, T. Zeuch, L. Seidel, L. Leon, and


F. Mauss. Combustion Chemistry of the Butane Isomers in Premixed Low-Pressure
Flames. Zeitschrift Physikalische Chemie, 225:1029–1054, 2011.

R. Ogink and V. Golovitchev. Gasoline HCCI Modeling: An Engine Cycle Simulation


Code with a Multi-Zone Combustion Model. SAE Technical Paper 2002-01-1745, 2002.
218 BIBLIOGRAPHY

S. Oh, S. Lee, Y. Choi, K. Kang, J. Cho, and K. Cha. Combustion and Emission Character-
istics in a Direct Injection LPG/Gasoline Spark Ignition Engine. SAE Technical Paper
2010-01-1461., 2010.

A.J. Pahnke, P.M. Cohen, and B.M. Sturgis. Preflame Oxidation of Hydrocarbons in a
Motored Engine. Ind. Eng. Chem. Res., pages 1024–1029, 1954.

J. Pan and C. Sheppard. A Theoretical and Experimental Study of the Modes of End Gas
Autoignition Leading to Knock in S. I. Engines. SAE Technical Paper 942060, 1994.

J. Pan, C.G.W. Sheppard, A. Tindall, M. Berzins, S.V. Pennington, and J.M. Ware. End Gas
Inhomogeneity, Autoignition and Knock. SAE Technical Paper 982616, 1998.

O. Park, P.S. Veloo, and F.N. Egolfopoulos. Flame studies of C2 hydrocarbons. Proc.
Combust. Inst, 34(1):711–718, 2013.

N. Pasadakis, V. Gaganis, and C. Foteinopoulos. Octane number prediction for gasoline


blends. Fuel Process. Technol., 87:505–509, 2006.

M. Pecqueur, K. Ceustermans, P. Huyskens, and D. Savvidis. Emissions generated from a


Suzuki Liane running on unleaded gasoline and LPG under the same load conditions.
SAE Technical Paper 2008-01-2637., 2008.

B.M. Penetrante, W.J. Pitz, M.C. Hsiao, B.T. Merritt, and G.E. Vogtlin. Effect of Hydro-
carbons on Plasma Treatment of NOx. Proceedings of the 1997 Diesel Engine Emissions
Reduction Workshop, 1997.

M.V. Petrova and F.A. Williams. A small detailed chemical-kinetic mechanism for hydro-
carbon oxidation. Combust. Flame, 144:526–544, 2006.

J.N. Pitts, H.W. Biermann, A.M. Winer, and E.C. Tuazon. Spectroscopic Identification and
Measurement of Gaseous Nitrous Acid in Dilute Auto Exhaust. Atmos. Environ., 18(4):
847–854, 1984.

W.J. Pitz and C.K. Westbrook. Modeling Chemical Kinetic Aspects of Engine Knock.
presented to the Spring meeting of the Western States Section of the Combustion Institute,
Boulder, CO, USA, 1984.
BIBLIOGRAPHY 219

W.J. Pitz and C.K. Westbrook. Chemical Kinetics of the High Pressure Oxidation of
n-Butane and its Relation to Engine Knock. Combust. Flame, 63:113, 1986.

S. Prabhu, H. Li, D. Miller, and N. Cernansky. The Effect of Nitric Oxide on Autoignition
of a Primary Reference Fuel Blend in a Motored Engine. SAE Technical Paper 932757,
1993.

S.K. Prabhu. Pre and Post-Combustion Hydrocarbon Oxidation Chemistry Related to Engine
Knock and Emissions. PhD thesis, Drexel University, 1997.

S.K. Prabhu, R.K. Bhat, D.L. Miller, and N.P. Cernansky. 1-Pentene Oxidation and Its
Interaction with Nitric Oxide in the Low and Negative Temperature Coefficient Regions.
Combust. Flame, 104:377–390, 1996.

S.K. Prabhu, X. Bian, D. Miller, and N. Cernansky. Post Combustion Hydrocarbon Oxida-
tion and Exhaust Emissions - Neat Fuel and Fuel Blend Studies. SAE Technical Paper
981456, 1998.

G. Protić-Lovasić, N. Jambrec, D. Deur-Siftar, and M.V. Prostenik. Determination of


catalytic reformed gasoline octane number by high resoltuion gas chromatography.
Fuel, 69:525–528, 1989.

A.D. Puckett. Knock ratings of gasoline substitutes. J. Res. Natl. Bur. Stand., 35 (4):273–284,
1945.

A.A. Quader. Effects of Spark Location and Combustion Duration on Nitric Oxide and
Hydrocarbon Emissions. SAE Technical Paper 730153, 1973.

B. Radu, G. Martin, R. Chiriac, and N. Apostolescu. On the Knock Characteristics of LPG


in a Spark Ignition Engine. SAE Technical Paper 2005-01-3773, 2005.

S. Radzimirski. Some Sources of Uncertainties in Motor Vehicle Emission Inventories-


Polish Experience. Presented at the European Environment Agency’s (EEA) Estimating
and reporting uncertainty in road transport emissions - greenhouse gases and air
pollutants workshop, Copenhagen, Denmark. Accessed 22/02/2013., November 2010.
URL eionet.europa.eu/events/transport_uncertainties/.
220 BIBLIOGRAPHY

M.M. Rahman, K.I. Hamada, M.M. Noor, K. Kadirgama, M.A. Maleque, and R.A. Bakar.
Heat Transfer Characteristics of an Intake Port for a Spark Ignition Engine: A Compara-
tive Study. J. Applied Sci., 10(18):2019–2026, 2010.

C.D. Rakopoulos and C.N. Michos. Development and validation of a multi-zone combus-
tion model for performance and nitric oxide formation in syngas fueled spark ignition
engine. Energ. Convers. Manage., 49:2924–2938, 2008.

C.L. Rasmussen, A.E. Rasmussen, and P. Glarborg. Sensitizing effects of NOX on CH4
oxidation at high pressure. Combust. Flame, 154:529–545, 2008.

M.J. Rauckis and W.J. McLean. The Effect of Hydrogen Addition on Ignition Delays and
Flame Propagation in Spark Ignition Engines. Combust. Sci. Tech., 19(5-6):207–216, 1979.

D. Razus, V. Brinzea, M. Mitu, C. Movileanu, and D. Oancea. Inerting effect of the


combustion products on the confined deflagration of Liquefied Petroleum Gas-air
mixtures. J. Loss. Prevnt. Proc., 22(4):463–468, 2009.

D. Razus, V. Brinzea, M. Mitu, and D. Oancea. Burning Velocity of Liquefied Petroleum


Gas (LPG)-Air Mixtures in the Presence of Exhaust Gas. Energ. Fuel, 24(3):1487–1494,
2010.

Reaction Design. CHEMKIN-PRO 10112, 2012. San Diego, CA.

S.U. Rege and R.T. Yang. Propane/propylene separation by pressure swing adsorption:
sorbent comparison and multiplicity of cyclic steady states. Chem. Eng. Sci., 57:1139–1149,
2002.

S. Regitz and N. Collings. Fast response air-to-fuel ratio measurements using a novel
device based on a wide band lambda sensor. Meas. Sci. Technol., 19, 2008.

J. Rudloff, J.-M. Zaccardi, S. Richard, and J.M. Anderlohr. Analysis of pre-ignition in


highly charged SI engines: Emphasis on the auto-ignition mode. Proc. Combust. Inst, 34:
2959–2967, 2013.

T.W. Ryan III, T.J. Callahan, and S.R. King. Engine Knock Rating of Natural Gases- Methane
Number. J. Eng. Gas Turbines Power, 115:769–776, 1993.
BIBLIOGRAPHY 221

J.R. Sabina. Correlation of Road and Laboratory Octane Numbers, Report from the
Cooperative Fuel Research Committee. SAE Trans., 43 (4):416–420, October 1938.

R. Sadeghbeigi. Fluid Catalytic Cracking Handbook. Butterworth-Heinemann, third edition,


2012. ISBN 978-0-12-386965-4. Oxford, UK.

S. Saito. Role of nuclear energy to a future of shortage of energy resources and global
warming. Journal of Nuclear Materials, 398:1–9, 2010.

Y. Sakai, H. Miyazaki, and K. Mukai. The Effect of Combustion Chamber Shape on


Nitrogen Oxides. SAE Technical Paper 730154, 1973.

K.C. Salooja. Studies Relating to the Mechanism of Antiknock Action of Tetraethyl Lead.
Combust. Flame, 9:211, 1965.

F.S. Schaub and R.L. Hubbard. A Procedure for Calculating Fuel Gas Blend Knock Rating
for Large-Bore Gas Engines and Predicting Engine Operation. J. Eng. Gas Turbines Power,
107:922–930, 1985.

H. Scheffé. Experiments with mixtures. J. Roy. Statist. Soc. Ser. B, 20:344–360, 1958.

E.J.Y. Scott. Knock characteristics of hydrocarbon mixtures. Proc. API, 38(3):90–101, 1958.

R.M. Shanmugam, N.M. Kankariya, J. Honvault, L. Srinivasan, H.C. Viswanatha, P. Nico-


las, N. Saravanan, and D. Christian. Performance and emission characterization of
1.2L MPI engine with multiple fuels (E10, LPG and CNG). SAE Int. J. Fuels Lubr., 3 (1):
334–352, 2010.

P. J. Shayler, S. A. May, and T. Ma. The Determination of Heat Transfer from the Combus-
tion Chambers of SI engines. SAE Technical Paper 931131, 1993.

C. Shelley. The Story of LPG. Poten and Partners, second edition, 2003.

C.G.W. Sheppard and G. Konig. End Gas Autoignition and Knock in a Spark Ignition
Engine. SAE Technical Paper 902135, 1990.

H.-H. Shin and W.-S. Yoon. Hydrocarbon Effects on the Promotion of Non-Thermal Plasma
NO-NO2 Conversion. Plasma Chem. Plasma P., 23(4), 2003.
222 BIBLIOGRAPHY

J.F. Sinnamon and M.C. Sellnau. A New Technique for Residual Gas Estimation and
Modeling in Engines. SAE Technical Paper 2008-01-0093, 2008.

J.R. Smith, R.M. Green, C.K. Westbrook, and W.J. Pitz. An Experimental and Modeling
Study of Engine Knock. Proc. Combust. Inst, 20:91–100, 1985.

D. Snelgrove, P. Dupont, and R. Bonetto. An Investigation into the Influence of LPG


(Autogas) Composition on the Exhaust Emissions and Fuel Consumption of 3 Bi-Fuelled
Renault Vehicles. SAE Technical Paper 961170, 1996.

H.S. Soyhan, H. Yasar, H. Walmsley, B. Head, G.T. Kalghatgi, and C. Sorusbay. Evaluation
of heat transfer correlations for HCCI engine modeling. Appl. Therm. Eng., 29(2-3):
541–549, 2009.

S. Soylu. Prediction of knock limited operating conditions of a natural gas engine. Energy
Convers. Manage., 46:121–138, 2005.

S.W. Sparrow. The Effect of Changes in Compression Ratio upon Engine Performance.
Technical report, NACA Report 205, 1925. United States National Bureau of Standards,
Gaithersburg, MD.

U. Spicher, H. Kroger, and J. Ganser. Detection of Knocking Combustion Using Simultane-


ously High-Speed Schlieren Cinematography and Multi Optical Fibre Technique. SAE
Technical Paper 912312, 1991.

Y.S. Stein, T.J. Held, and F.L. Dryer. Reaction Studies on n-Butane and Oxygen Systems
at 12.5 atmospheres and 500-800 K. Presented at the Fall Meeting of the Western States
Section/The Combustion Institute, Berkeley, CA, 12-13 October, 1992. Paper No. WSSSCI
92-92.

O. Stenlaas, P. Einewall, R. Egnell, and B. Johansson. Measurement of Knock and Ion


Current in a Spark Ignition Engine with and without NO Addition to the Intake Air.
SAE Technical Paper 2003-01-0639, 2003.

O. Stenlass, A. Gogan, R. Egnell, B. Sunden, and F. Mauss. The Influence of Nitric Oxide on
the Occurrence of Autoignition in the End Gas of Spark Ignition Engines. SAE Technical
Paper 2002-01-2699, 2002.
BIBLIOGRAPHY 223

D. Stivender. Development of a Fuel-Based Mass Emission Measurement Procedure. SAE


Technical Paper 710604, 1971.

A. Swarts. Insights relating to octane rating and the underlying role of autoignition. PhD thesis,
University of Cape Town, February 2006.

A. Swarts, C. Viljoen, and R. Coetzer. The Analysis of Observed Burn Rates in a Spark-
Ignition Engine and the Relation to Fuel Properties. SAE Technical Paper 2003-01-3125,
2003.

A. Swarts, A. Yates, C. Viljoen, and R. Coetzer. Standard Knock Intensity Revisited:


Atypical Burn Rate Characteristics identified in the CFR Octane Rating Engine. SAE
Technical Paper 2004-01-1850, 2004.

A. Swarts, A. Yates, C. Viljoen, and R. Coetzer. A Further Study of Inconsistencies between


Autoignition and Knock Intensity in the CFR Octane Rating Engine. SAE Technical Paper
2005-01-2081, 2005.

R.J. Tabaczynski, J.B. Heywood, and J.C. Keck. Time-Resolved Measurements of Hydro-
carbon Mass Flowrate in the Exhaust of a Spark-Ignition Engine. SAE Technical Paper
720112, 1972.

Y. Tan, C.G. Fotache, and C.K. Law. Effects of NO on the ignition of hydrogen and
hydrocarbons by heated counterflowing air. Combust. Flame, 119(3):346–355, 1999.

C. Tang, L. Wei, J. Zhang, X. Man, and Z. Huang. Shock Tube Measurements and Kinetic
Investigation on the Ignition Delay Times of Methane/Dimethyl Ether Mixtures. Energ.
Fuels, 26(11):6720–6728, 2012.

C.F. Taylor. The Internal Combustion Engine in Theory and Practice: Combustion, Fuels,
Materials, Design. MIT Press, second edition, 1985. ISBN 9780262700276.

The Adept Group Inc. California Air Resources Board Liquefied Petroleum Gas Fuel Blends
Evaluation Project; May 1998 Progress Report, July 1998. URL arb.ca.gov/fuels/
altfuels/lpg/mvlpge/rptmay98.pdf. Los Angeles, CA. Accessed 03/11/2012.
224 BIBLIOGRAPHY

The Adept Group Inc. California Air Resources Board Liquefied Petroleum Gas Fuel Blends
Evaluation Project; Final Report, February 2000. URL adeptgroup.net/Reports/
0002hd5report.pdf. Los Angeles, CA. Accessed 03/11/2012.

C. Thiel, A. Perujo, and A. Mercier. Cost and CO2 aspects of future vehicle options in
Europe under new energy policy scenarios. Energ. Policy, 38(11):7142–7151, 2010.

H.W. Thompson and C.N. Hinshelwood. The Influence of Nitrogen Peroxide on the
Combination of Hydrogen and Oxygen. Proc. R. Soc. Lond. A, pages 219–227, 1929.

N. Thompson and J. Wallace. Effect of Engine Operating Variables and Piston and Ring
Parameters on Crevice Hydrocarbon Emissions. SAE Technical Paper 940480, 1994.

E. Tribbett, E. Froehlich, and L. Bayer. Effects of Ignition Timing, Equivalence Ratio and
Compression Ratio on RDH Engine Performance. Department of Mechanical Engineering,
Stanford University, 2010.

W.R. True. Global LPG supply growth responding to high oil prices. Oil Gas J., 110 (3b):23,
2012.

S.R. Turns. An introduction to combustion: concepts and applications. McGraw-Hill, second


edition, 2000. ISBN 0-07-230096-5.

T. Tzanetakis, P. Singh, J. Chen, and M.J. Thomson. Knock limit prediction via multi-zone
modelling of a primary reference fuel HCCI engine. Int. J. Vehicle Design, 54(1):47–72,
2010.

A.R. Ubbelohde, J.W. Drinkwater, and A. Egerton. Pro-knocks and Hydrocarbon Combus-
tion. Proc. R. Soc. Lond. A, 103(15):103–115, 1935.

United States Energy Information Administration. Review of Emerging Resources: U.S.


Shale Gas and Shale Oil Plays. Technical report, July 2011a. U.S. Department of Energy,
Washington, DC.

United States Energy Information Administration. World Shale Gas Resources: An Initial
Assessment of 14 Regions Outside the United States. Technical report, April 2011b. U.S.
Department of Energy, Washington, DC.
BIBLIOGRAPHY 225

United States Energy Information Administration. Technically Recoverable Shale Oil and
Shale Gas Resources: An Assessment of 137 Shale Formations in 41 Countries Outside
the United States. Technical report, June 2013. U.S. Department of Energy, Washington,
DC.

United States Geological Survey. National Assessment of Oil and Gas Fact Sheet. Technical
report, April 2013. U.S. Department of the Interior, Reston, VA.

J.A. van Leeuwen, R.J. Jonker, and R. Gill. Octane number prediction based on gas
chromatographic analysis with non-linear regression techniques. Chemometr. Intell. Lab.,
25:325–340, 1994.

J. Vancoillie. CFR Engine forward/reverse discharge coefficient data for the intake and
exhaust valves. Ghent University, Personal Communication, March 2013.

S. Verhelst and C.G.W. Sheppard. Multi-zone thermodynamic modelling of spark-ignition


engine combustion- An overview. Energy Convers. Manage., 50:1326–1335, 2009.

I.I. Vibe. Semi-empirical expression for combustion rate in engines. Proc. Conference on
Piston Engines, USSR Academy of Sciences, Moscow, pages 185–191, 1956.

C. Walcutt, J.M. Mason, and E.B. Rifkin. Effect of Preflame Oxidation Reactions on Engine
Knock. J. Ind. Eng. Chem., 46(5):1029–1034, 1954.

H.C. Watson and D.R.R. Gowdie. The systematic evaluation of twelve LP Gas fuels for
emissions and fuel consumption. SAE Technical Paper 2000-01-1867., 2000.

H.C. Watson and E.E. Milkins. Comparison and optimisation of emissions, efficiency and
power of five automotive fuels in one engine. Int. J. of Vehicle Design, 3(4):463–476, 1982.

H.C. Watson and P. Phuong. Why Liquid Phase LPG Port Injection has Superior Power
and Efficiency to Gas Phase Port Injection. SAE Technical Paper 2007-01-3552., 2007.

P. Weiss and J.C. Keck. Fast Sampling Valve Measurements of Hydrocarbons in the
Cylinder of a CFR Engine. SAE Technical Paper 810149, 1981.

C.K. Westbrook and W. J. Pitz. Modeling of Knock in Spark-Ignition Engines. Int. Symp.
COMODIA, pages 11–20, 1990.
226 BIBLIOGRAPHY

R.W. Wheeler. Abnormal Combustion Effects on Economy. In: Fuel Economy in Road Vehicles
Powered by Spark Ignition Engines. Plenum Press, New York, NY, 1984. ISBN 0-306-41438-
4.

N.O. Wittmann and J.H. Smith. Piston temperature measurement and piston design
investigation on a C.F.R. engine. Technical report, Massachusetts Institute of Technology,
1946. URL http://hdl.handle.net/10945/6510. Accessed 18/03/2012.

P.O. Witze and R.M. Green. Determining the Location of End-Gas Autoignition Using
Ionisation Probes Installed In the Head Gasket. SAE Technical Paper 932645, 1993.

World LP Gas Association. Autogas incentive policies. Technical report, July 2012.
URL worldlpgas.com/resources/publications. Neuilly-sur-Seine, France. Ac-
cessed 22/02/2012.

G. Woschni. A Universally Applicable Equation for the Instantaneous Heat Transfer


Coefficient in the Internal Combustion Engine. SAE Technical Paper 670931, 1967.

A.D.B. Yates and C.L. Viljoen. An Improved Empirical Model for Describing Auto-ignition.
SAE Technical Paper 2008-01-1629, 2008.

A.D.B. Yates, A. Swarts, and C.L. Viljoen. An Investigation of Anomalies Identified within
the ASTM Research and Motor Octane Scales. SAE Technical Paper 2003-01-1772, 2003.

S. Yousufuddin and S.N. Mehdi. Performance and Emission Characteristics of LPG-Fuelled


Variable Compression Ratio SI Engine. Turkish J. Eng. Env. Sci., 32:7–12, 2008.

J. Zhang, E. Hu, Z. Zhang, L. Pan, and Z. Huang. Comparative Study on Ignition Delay
Times of C1-C4 Alkanes. Energ. Fuel., 27(6):3480–3487, 2013.
Appendix A

Research and Motor octane number


data

Table A.1: Measured RONs and MONs (primary data set). Mixture mole fractions have
been rounded to one decimal place.

Fuel composition (% mol)


propane propylene n-butane iso-butane RON MON
100.0 109.4 96.3
100.0 100.2 83.3
100.0 93.5 89.0
100.0 100.1 96.8
50.0 50.0 103.3 89.9
50.0 50.0 100.4 92.6
50.0 50.0 104.4 97.3
50.0 50.0 97.6 86.9
50.0 50.0 101.5 91.6
50.0 50.0 96.8 92.8
66.7 16.7 16.7 104.6 92.9
33.3 33.3 33.3 100.9 89.8
16.7 66.7 16.7 100.9 87.1
16.7 16.7 66.7 97.2 89.8
66.7 16.7 16.7 106.1 94.8
33.3 33.3 33.3 103.7 93.2
16.7 66.7 16.7 102.2 88.8
16.7 16.7 66.7 102.1 95.4
66.7 16.7 16.7 104.9 95.2
33.3 33.3 33.3 100.3 94.5
16.7 66.7 16.7 96.6 91.8
16.7 16.7 66.7 100.1 95.5
66.7 16.7 16.7 100.0 87.7
33.3 33.3 33.3 98.7 90.7
16.7 66.7 16.7 95.9 90.0
16.7 16.7 66.7 99.5 93.7

227
228 Research and Motor octane number data

Table A.2: Measured RONs and MONs (check point data set). Mixture mole fractions have
been rounded to one decimal place.

Fuel composition (% mol)


propane propylene n-butane iso-butane RON MON
75.0 25.0 105.8 93.0
75.0 25.0 104.6 94.4
75.0 25.0 107.0 97.0
25.0 75.0 101.4 86.5
25.0 75.0 96.7 90.9
25.0 75.0 102.2 97.3
75.0 25.0 99.2 85.4
75.0 25.0 101.3 88.1
25.0 75.0 95.6 88.1
25.0 75.0 101.0 94.5
75.0 25.0 95.1 91.1
25.0 75.0 98.4 94.9
41.7 41.7 16.7 102.4 90.1
41.7 16.7 41.7 100.7 91.5
16.7 41.7 41.7 99.4 88.6
41.7 41.7 16.7 103.8 91.4
41.7 16.7 41.7 104.2 95.3
16.7 41.7 41.7 102.6 92.3
41.7 41.7 16.7 100.4 93.3
41.7 16.7 41.7 102.4 95.6
16.7 41.7 41.7 98.4 93.5
41.7 41.7 16.7 98.1 89.0
41.7 16.7 41.7 100.1 91.1
16.7 41.7 41.7 97.8 91.9

Table A.3: Measured RONs and MONs (quaternary mixture data set). Mixture mole
fractions have been rounded to one decimal place.

Fuel composition (% mol)


propane propylene n-butane iso-butane RON MON
70.0 10.0 10.0 10.0 105.6 94.8
50.0 16.7 16.7 16.7 103.4 93.3
16.7 50.0 16.7 16.7 101.2 89.5
16.7 16.7 50.0 16.7 98.4 91.0
16.7 16.7 16.7 50.0 101.0 93.9
10.0 70.0 10.0 10.0 101.2 87.5
10.0 10.0 70.0 10.0 96.2 90.3
10.0 10.0 10.0 70.0 100.5 95.3
Appendix B

Empirical data used in the GT-Power


model

B.1 Intake and exhaust valve lift profiles

The intake and exhaust valve lift profiles were measured directly on the CFR engine used
in the experimental work. This data was acquired in increments of 1.0 crank angle degree
using a dial indicator and crank indexing wheel. The measured intake and exhaust valve
lift profiles are presented in Figure B.1.

Figure B.1: Measured intake and exhaust valve lift profiles for the CFR engine used in the
experimental work. This data was used in the GT-Power gas exchange sub-models.

229
230 Empirical data used in the GT-Power model

B.2 Intake and exhaust valve discharge coefficient data

The discharge coefficients for the standard CFR engine intake and exhaust valves were
unable to be measured directly. This data was therefore obtained from the literature
(Vancoillie, 2013). The discharge coefficients were measured using a standard CFR engine
cylinder head and a steady-state flow bench. The latter was operated with a constant
pressure difference across the valves of 20.0 in. Hg.

The discharge coefficients are provided to GT-Power as a function of a non-dimensional


(reference) array. In this work, the non-dimensional array was,

Vlift
reference array = ,
Vreference diameter (B.1)

where Vlift is the instantaenous valve lift (Figure B.1) and Vreference diameter is a reference
diameter. In the case of the latter, the reference diameter was assumed to be the valve
diameter (Table B.1). The resulting intake and exhaust discharge coefficient profiles utilised
by the gas exchange sub-models in GT-Power are presented in Figure B.2.

Table B.1: Reference diameters used to non-dimensionalise the valve lift profiles (Vancoillie,
2013).

Valve Reference diameter (mm)


Intake 34.18
Exhaust 34.45
B.2 Intake and exhaust valve discharge coefficient data 231

(a) Intake valve.

(b) Exhaust valve.

Figure B.2: Forward and reverse discharge coefficients for the standard CFR engine intake
and exhaust valves versus the non-dimensionalised valve lift. Data was acquired using a
steady-state flow blench, and was obtained from the literature (Vancoillie, 2013).
Appendix C

Calculation of the ideal products of


combustion

In this analysis, air is assumed to be comprised of only nitrogen (N2 ) and oxygen (O2 ). It
is also assumed that for each mole of O2 , 3.76 moles of N2 are present. Therefore, the air is
assumed to be comprised of 79% N2 and 21% O2 .

Assuming no dissociation of the combustion products takes place, the combustion of any
hydrocarbon fuel, Cx Hy , in the presence of air can be expressed as,

Cx Hy + a (O2 + 3.76 N2 ) → b CO2 + c CO + d H2 O + e H2 + f O2 + 3.76 a N2 (C.1)

C.1 Stoichiometric or fuel-lean combustion (φ ≤ 1)

There is sufficient O2 available to fully convert Cx Hy into CO2 and H2 O. Therefore, the
coefficients c and e are zero, and Equation C.1 can be re-written as,

Cx Hy + a (O2 + 3.76 N2 ) → b CO2 + d H2 O + f O2 + 3.76 a N2 (C.2)

The coefficient a can be related to the equivalence ratio, φ,

233
234 Calculation of the ideal products of combustion

x + y/4
a = (C.3)
φ

The equivalence ratio is related to the air-fuel ratio, (A/F), and the stoichiometric air-fuel
ratio, (A/F)stoich by,

(A/F)stoich
φ = (C.4)
(A/F)

Therefore, the value of the coefficient a can be computed for a given Cx Hy and known
value of φ.

The values of the coefficients b, d and f can be computed using C-atom, H-atom and
O-atom balances, respectively,

b = x (C.5)

y
d = (C.6)
4

1−φ
 
y
f = x+ (C.7)
φ 4

The total number of moles of the products per mole of fuel, Ntotal , is computed by summing
the coefficients in Equations C.5− C.7 with the 3.76 a moles of N2 . This yields,

 
y x + y/4
Ntotal = x+ + (1 − φ + 3.76) (C.8)
2 φ

The product mole fractions can therefore be expressed as,


C.2 Fuel-rich combustion (φ > 1) 235

x
χCO2 = (C.9)
Ntotal

y/2
χ H2 O = (C.10)
Ntotal

h  i
1−φ
φ (x + y/4)
χO2 = (C.11)
Ntotal

[3.76 (x + y/4)]
χ N2 = (C.12)
Ntotal

C.2 Fuel-rich combustion (φ > 1)

With reference to Equation C.1, there is insufficient O2 available to fully convert Cx Hy into
CO2 and H2 O. Therefore, the coefficient f is zero, and Equation C.1 can be re-written as,

Cx Hy + a (O2 + 3.76 N2 ) → b CO2 + c CO + d H2 O + e H2 + 3.76 a N2 (C.13)

There are four unknowns in Equation C.13. However, only three equations can be formed
on the basis of C-atom, H-atom and O-atom balances. Therefore, the water-gas shift
reaction is employed to yield a fourth equation,

CO + H2 O ⇐⇒ CO2 + H2 (C.14)

Equation C.14 is used to account for the co-existence of CO and H2 O in the combustion
products. The water-gas shift equilibrium can be expressed as,
236 Calculation of the ideal products of combustion

PCO2 /P0 PH2 /P0


 
be
Kp = 0 0
= (C.15)
(PCO /P ) (PH2 O /P ) cd

Performing the C-atom, H-atom and O-atom balances in terms of the coefficient b yields,

c = x−b (C.16)

d = 2a − b − x (C.17)

y
e = −2 a + b + x + (C.18)
2

Substituting Equations C.16−C.18 into Equation C.15 yields,


2 a K p − 1 + x + y/2
b = 
2 Kp − 1

1
q
 2 
−  2 a K p − 1 + x + y/2 − 4 K p K p − 1 (2 a x − x2 )
2 Kp − 1
(C.19)

To obtain a physical (non-negative) solution to Equation C.19, the negative root is chosen.
The equilibrium constant for the water-gas shift reaction is evaluated as a function of
temperature using a polynomial fit of tabulated data (Table C.1). The total numbers of
moles of combustion products can therefore be expressed as,

y
Ntotal = b + c + d + e + 3.76 a = x + + 3.76 a (C.20)
2
C.2 Fuel-rich combustion (φ > 1) 237

The product mole fractions can therefore be expressed as,

b
χCO2 = (C.21)
Ntotal

c (x − b )
χCO = = (C.22)
Ntotal Ntotal

d (2 a − b − x)
χH2 O = = (C.23)
Ntotal Ntotal

e (−2 a + b + x + y/2)
χH2 = = (C.24)
Ntotal Ntotal

3.76 a
χN2 = (C.25)
Ntotal

where the coefficient a in Equation C.25 is evaluated using Equation C.3.

Table C.1: Selected values for the water-gas shift reaction equilibrium constant as a function
of temperature (Turns, 2000).

Temperature (K) Kp
500 138.3
1000 1.443
1500 0.3887
2000 0.2200
2500 0.1635
3000 0.1378
3500 0.1241
Appendix D

Speciated exhaust hydrocarbon


measurements for propane

The effect of unreacted and partially reacted hydrocarbons on the autoignition of propane
was examined using the speciated exhaust data reported in Figure D.1. This data was
acquired with a multi-cylinder engine operated on propane. The compression ratio, engine
speed and spark timing were 8.0:1, 1500 rpm and 23.0 ◦ BTDC respectively. The unreacted
fuel and stable intermediate hydrocarbons in the exhaust were measured using a flame-
ionization detector. This technique identified more than 90% of the exhaust hydrocarbons.

The total hydrocarbon measurements reported by Kaiser et al. agree closely with those
reported in an earlier study that utilised a propane-fuelled CFR engine (Daniel, 1967). In
this study, the compression ratio, engine speed and spark timing were 9.0:1, 1360 rpm
and 2.0-38.0 ◦ BTDC respectively. It can therefore be argued that the total hydrocarbon
emissions do not vary significantly with small changes in the compression ratio, engine
speed or spark timing.

These arguments are supported in a subsequent study undertaken by the same authors
(Kaiser et al., 1984). Figure D.2 shows that at a fixed equivalence ratio of φ=1.05, a
10% increase in the compression ratio (from 8.0:1) yields a near linear increase of 14%
in the total measured exhaust hydrocarbon emissions. However, Figure D.1a indicates
that the same 10% increase/reduction in the equivalence ratio (from φ=1.05) at fixed
spark timing and compression ratio yields a 34% increase/reduction in the total exhaust

239
240 Speciated exhaust hydrocarbon measurements for propane

(a) Total exhaust hydrocarbon emissions. (b) Speciated exhaust hydrocarbon concentrations.

Figure D.1: Variation in the total measured exhaust hydrocarbon emissions for propane as a
function of equivalence ratio, along with the corresponding speciated exhaust hydrocarbon
emissions (Kaiser et al., 1983). C2 H6 /C2 H4 = 0.11 (1.0 ≤ φ ≤ 1.3).

Figure D.2: Total measured exhaust hydrocarbon emissions as a function of compression


ratio. Data was obtained in a CFR engine operated on propane at 1500 rpm, with the spark
timing set at 22.0 ◦ BTDC (Kaiser et al., 1984).
241

hydrocarbon emissions. Thus, the equivalence ratio appears to be the most dominant
factor in establishing the total hydrocarbon emissions.

In the analysis undertaken in Chapter 6, the equivalence ratio is therefore matched


to the SKI value for propane (λ = 0.96), as shown in Figure D.1a. This determines the
total hydrocarbon emissions (1500 ppm C1 ). The concentrations of the individual species
are then computed using the data presented in Figure D.1b. The concentrations of the
individual hydrocarbons are summarised in Chapter 6.
Appendix E

Chemical kinetic mechanism

243
Reaction A B E Auxiliary data

244
1 H+O2=O+OH 3.55E+15 -0.406 16600 REV/1.027E13 -1.5E-2 -1.33E2/
2 O+H2=H+OH 5.08E+04 2.67 6292 REV/2.637E4 2.651E0 4.88E3/
3 OH+H2=H+H2O 2.16E+08 1.51 3430 REV/2.29E9 1.404E0 1.832E4/
4 O+H2O=2OH 2.97E+06 2.02 13400 REV/1.454E5 2.107E0 -2.904E3/
5 H2+M=2H+M 4.58E+19 -1.4 104400 Third body: H2 /2.5/ Third body: H2O /12.0/
Third body: CO /1.9/ Third body: CO2 /3.8/
REV/1.145E20 -1.676E0 8.2E2/
6 O2+M=2O+M 4.42E+17 -0.634 118900 Third body: H2 /2.5/ Third body: H2O /12.0/
Third body: AR /0.83/ Third body: CO /1.9/
Third body: CO2 /3.8/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.83/
REV/6.165E15 -5.0E-1 0.0E0/
7 OH+M=O+H+M 9.78E+17 -0.743 102100 Third body: H2 /2.5/ Third body: H2O /12.0/
Third body: AR /0.75/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.75/
REV/4.714E18 -1.0E0 0.0E0/
8 H2O+M=H+OH+M 1.91E+23 -1.83 118500 Third body: H2 /0.73/ Third body: H2O /12.0/
Third body: AR /0.38/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.38/
REV/4.5E22 -2.0E0 0.0E0/
9 H+O2(+M)=HO2(+M) 1.48E+12 0.6 0 Third body: H2 /1.3/ Third body: H2O /14.0/
Third body: AR /0.67/ Third body: CO /1.9/
Third body: CO2 /3.8/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.67/
LOW/3.482E16 -4.11E-1 -1.115E3/ TROE/5.0E-1
1.0E-30 1.0E30 1.0E10/
10 HO2+H=H2+O2 1.66E+13 0 823 REV/3.166E12 3.48E-1 5.551E4/
11 HO2+H=2OH 7.08E+13 0 295 REV/2.028E10 7.2E-1 3.684E4/
Chemical kinetic mechanism

12 HO2+O=OH+O2 3.25E+13 0 0 REV/3.217E12 3.29E-1 5.328E4/


Reaction A B E Auxiliary data
13 HO2+OH=H2O+O2 2.89E+13 0 -497 REV/5.844E13 2.42E-1 6.908E4/
14 H2O2+O2=2HO2 4.63E+16 -0.347 50670 REV/4.2E14 0.0E0 1.198E4/ DUP
15 H2O2+O2=2HO2 1.43E+13 -0.347 37060 REV/1.3E11 0.0E0 -1.629E3/ DUP
16 H2O2(+M)=2OH(+M) 2.95E+14 0 48430 Third body: H2 /2.5/ Third body: H2O /12.0/
Third body: AR /0.64/ Third body: CO /1.9/
Third body: CO2 /3.8/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.64/
LOW/1.202E17 0.0E0 4.55E4/ TROE/5.0E-1 1.0E-
30 1.0E30 1.0E10/
17 H2O2+H=H2O+OH 2.41E+13 0 3970 REV/1.265E8 1.31E0 7.141E4/
18 H2O2+H=H2+HO2 6.03E+13 0 7950 REV/1.041E11 6.95E-1 2.395E4/
19 H2O2+O=OH+HO2 9.55E+06 2 3970 REV/8.568E3 2.676E0 1.856E4/
20 H2O2+OH=H2O+HO2 1.00E+12 0 0 REV/1.833E10 5.89E-1 3.089E4/ DUP
21 H2O2+OH=H2O+HO2 5.80E+14 0 9557 REV/1.063E13 5.89E-1 4.045E4/ DUP
22 CO+O(+M)=CO2(+M) 1.80E+10 0 2384 Third body: H2 /2.0/ Third body: O2 /6.0/ Third
body: H2O /6.0/ Third body: AR /0.5/ Third
body: CO /1.5/ Third body: CO2 /3.5/ Third
body: CH4 /2.0/ Third body: C2H6 /3.0/ Third
body: HE /0.5/ LOW/1.35E24 -2.788E0 4.191E3/
23 CO+O2=CO2+O 1.05E+12 0 42540 REV/8.035E15 -8.0E-1 5.123E4/
24 CO+OH=CO2+H 2.23E+05 1.89 -1158 REV/5.896E11 6.99E-1 2.426E4/
25 CO+HO2=CO2+OH 1.57E+05 2.18 17940 REV/1.189E8 1.71E0 7.991E4/
26 HCO+M=H+CO+M 4.75E+11 0.66 14870 Third body: H2 /2.0/ Third body: H2O /12.0/
Third body: CO /1.5/ Third body: CO2 /2.0/
Third body: CH4 /2.0/ Third body: C2H6 /3.0/
REV/3.582E10 1.041E0 -4.573E2/
27 HCO+O2=CO+HO2 2.71E+10 0.68 -469 REV/4.283E9 9.89E-1 3.307E4/
28 HCO+H=CO+H2 7.34E+13 0 0 REV/2.212E12 6.56E-1 8.823E4/
29 HCO+O=CO+OH 3.02E+13 0 0 REV/4.725E11 6.38E-1 8.682E4/
30 HCO+O=CO2+H 3.00E+13 0 0 REV/1.241E18 -5.53E-1 1.122E5/
31 HCO+OH=CO+H2O 1.02E+14 0 0 REV/3.259E13 5.51E-1 1.031E5/
32 HCO+CH3=CH4+CO 2.65E+13 0 0 REV/7.286E14 2.11E-1 8.977E4/
245
Reaction A B E Auxiliary data

246
33 HCO+HO2=CH2O+O2 2.50E+14 -0.061 13920 REV/8.07E15 0.0E0 5.342E4/
34 HCO+HO2=¿CO2+H+OH 3.00E+13 0 0
35 O2CHO=HCO+O2 1.40E+29 -4.549 46300 REV/1.2E11 0.0E0 -1.1E3/
36 CH2O+O2CHO=HCO+HO2CHO 1.99E+12 0 11660 REV/3.744E3 1.952E0 7.145E3/
37 HO2CHO=OCHO+OH 5.01E+14 0 40150 REV/2.871E6 2.094E0 -7.012E3/
38 OCHO+M=H+CO2+M 5.32E+14 -0.353 17580 REV/7.5E13 0.0E0 2.9E4/
39 CH2O+CO=2HCO 9.19E+13 0.37 73040 REV/1.8E13 0.0E0 0.0E0/
40 2HCO=¿H2+2CO 3.00E+12 0 0
41 HCO+H(+M)=CH2O(+M) 1.09E+12 0.48 -260 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.7/
LOW/1.35E24 -2.57E0 1.425E3/ TROE/7.824E-1
2.71E2 2.755E3 6.57E3/
42 CO+H2(+M)=CH2O(+M) 4.30E+07 1.5 79600 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.7/
LOW/5.07E27 -3.42E0 8.4348E4/ TROE/9.32E-1
1.97E2 1.54E3 1.03E4/
43 CH2O+OH=HCO+H2O 7.82E+07 1.63 -1055 REV/4.896E6 1.811E0 2.903E4/
44 CH2O+H=HCO+H2 5.74E+07 1.9 2740 REV/3.39E5 2.187E0 1.793E4/
45 CH2O+O=HCO+OH 6.26E+09 1.15 2260 REV/1.919E7 1.418E0 1.604E4/
46 CH2O+CH3=HCO+CH4 3.83E+01 3.36 4312 REV/2.063E2 3.201E0 2.104E4/
47 CH2O+HO2=HCO+H2O2 7.10E-03 4.517 6580 REV/2.426E-2 4.108E0 5.769E3/
48 HOCH2O=CH2O+OH 2.06E+21 -2.336 25730 REV/4.5E15 -1.1E0 0.0E0/
49 HOCH2O=HOCHO+H 1.00E+14 0 14900 REV/1.123E15 -2.95E-1 1.15E4/
50 HOCHO=CO+H2O 2.45E+12 0 60470 REV/2.255E3 2.093E0 5.289E4/
51 HOCHO=CO2+H2 2.95E+09 0 48520 REV/6.772E5 1.008E0 5.147E4/
52 HOCHO=HCO+OH 3.47E+22 -1.542 110700 REV/1.0E14 0.0E0 0.0E0/
53 HOCHO+O2=OCHO+HO2 4.10E+12 -0.308 59880 REV/3.5E10 0.0E0 -3.275E3/
54 HOCHO+OH=¿H2O+CO2+H 2.62E+06 2.06 916
55 HOCHO+OH=¿H2O+CO+OH 1.85E+07 1.51 -962
56 HOCHO+H=¿H2+CO2+H 4.24E+06 2.1 4868
Chemical kinetic mechanism
Reaction A B E Auxiliary data
57 HOCHO+H=¿H2+CO+OH 6.03E+13 -0.35 2988
58 HOCHO+CH3=¿CH4+CO+OH 3.90E-07 5.8 2200
59 HOCHO+HO2=OCHO+H2O2 2.55E+12 0.04 34470 REV/2.4E12 0.0E0 1.0E4/
60 HOCHO+HO2=¿H2O2+CO+OH 1.00E+12 0 11920
61 HOCHO+O=¿CO+2OH 1.77E+18 -1.9 2975
62 HOCHO+HCO=CH2O+OCHO 8.58E+11 0.04 26750 REV/5.6E12 0.0E0 1.36E4/
63 CH3O(+M)=CH2O+H(+M) 6.80E+13 0 26170 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: CO /1.5/ Third body: CO2 /2.0/
Third body: CH4 /2.0/ Third body: C2H6 /3.0/
LOW/1.867E25 -3.0E0 2.4307E4/ TROE/9.0E-1
2.5E3 1.3E3 1.0E99/
64 CH3O+O2=CH2O+HO2 4.38E-19 9.5 -5501 REV/1.416E-20 9.816E0 2.108E4/
65 CH2O+CH3O=CH3OH+HCO 6.62E+11 0 2294 REV/8.393E10 7.4E-2 1.771E4/
66 CH4+CH3O=CH3+CH3OH 6.12E+02 2.867 8248 REV/1.44E1 3.1E0 6.935E3/
67 CH3O+CH3=CH2O+CH4 1.20E+13 0 0 REV/6.749E13 2.18E-1 8.281E4/
68 CH3O+H=CH2O+H2 2.00E+13 0 0 REV/1.233E11 6.64E-1 8.127E4/
69 CH3O+HO2=CH2O+H2O2 3.01E+11 0 0 REV/1.074E12 -3.1E-2 6.527E4/
70 CH2O+H(+M)=CH2OH(+M) 5.40E+11 0.454 3600 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: CO /1.5/ Third body: CO2 /2.0/
Third body: CH4 /2.0/ Third body: C2H6 /3.0/
LOW/1.27E32 -4.82E0 6.53E3/ TROE/7.187E-1
1.03E2 1.291E3 4.16E3/
71 CH2OH+O2=CH2O+HO2 1.51E+15 -1 0 REV/1.975E14 -5.8E-1 2.006E4/ DUP
72 CH2OH+O2=CH2O+HO2 2.41E+14 0 5017 REV/3.152E13 4.2E-1 2.508E4/ DUP
73 CH2OH+H=CH2O+H2 6.00E+12 0 0 REV/1.497E11 7.68E-1 7.475E4/
74 CH2OH+HO2=CH2O+H2O2 1.20E+13 0 0 REV/1.732E14 7.3E-2 5.875E4/
75 CH2OH+HCO=2CH2O 1.80E+14 0 0 REV/7.602E14 4.81E-1 5.956E4/
76 CH2OH+CH3O=CH2O+CH3OH 2.40E+13 0 0 REV/1.285E13 5.55E-1 7.498E4/
77 CH2OH+CH2O=CH3OH+HCO 1.88E+04 2.722 4208 REV/9.63E3 2.9E0 1.311E4/
78 OH+CH2OH=H2O+CH2O 2.40E+13 0 0 REV/6.347E12 6.62E-1 8.964E4/
79 O+CH2OH=OH+CH2O 4.20E+13 0 0 REV/5.438E11 7.49E-1 7.334E4/
80 CH2O+CH3OH=2CH2OH 6.50E+12 0.659 68460 REV/3.0E12 0.0E0 0.0E0/
81 CH2OH+HO2=HOCH2O+OH 1.00E+13 0 0 REV/8.169E13 -2.4E-2 3.347E4/
82 OCH2O2H=CH2O+HO2 1.28E+18 -1.8 10460 REV/1.5E11 0.0E0 1.19E4/
83 OCH2O2H=HOCH2O2 3.00E+11 0 8600 REV/4.241E8 9.5E-1 2.62E4/
247
Reaction A B E Auxiliary data

248
84 HOCH2O2+HO2=HOCH2O2H+O2 3.50E+10 0 -3275 REV/1.046E14 -8.4E-1 3.487E4/
85 HOCH2O2H=HOCH2O+OH 1.02E+21 -1.92 42490 REV/1.0E13 0.0E0 0.0E0/
86 CH3OH(+M)=CH3+OH(+M) 1.90E+16 0 91730 LOW/2.95E44 -7.35E0 9.546E4/ TROE/4.14E-1
2.79E2 5.459E3 1.0E10/
87 CH3OH(+M)=CH2OH+H(+M) 2.69E+16 -0.08 98940 LOW/2.34E40 -6.33E0 1.031E5/ TROE/7.73E-1
6.93E2 5.333E3 1.0E10/
88 CH3OH+H=CH3O+H2 3.60E+12 0 6095 REV/1.677E11 2.12E-1 5.868E3/
89 CH3OH+H=CH2OH+H2 1.20E+06 2.4 2583 REV/1.386E4 2.509E0 8.871E3/
90 CH3OH+O=CH2OH+OH 3.88E+05 2.5 3080 REV/2.319E3 2.59E0 7.956E3/
91 CH3OH+OH=CH3O+H2O 5.13E+05 2.13 2450 REV/2.534E5 2.237E0 1.712E4/
92 CH3OH+OH=CH2OH+H2O 1.44E+06 2 -839 REV/1.758E5 2.003E0 2.034E4/
93 CH3OH+O2=CH2OH+HO2 2.05E+13 0 44900 REV/1.238E12 -2.39E-1 -3.501E3/
94 CH3OH+HO2=CH2OH+H2O2 1.08E+04 2.55 10530 REV/7.195E4 1.963E0 8.19E2/
95 CH3OH+CH3=CH2OH+CH4 3.19E+01 3.17 7172 REV/3.351E2 2.833E0 1.5E4/
96 CH3O+CH3OH=CH2OH+CH3OH 3.00E+11 0 4074 REV/7.416E10 -1.04E-1 1.059E4/
97 CH3OH+CH2O=2CH3O 7.98E+12 0.452 81490 REV/6.03E13 0.0E0 0.0E0/
98 CH3+H(+M)=CH4(+M) 2.14E+15 -0.4 0 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4
/2.0/ Third body: C2H6 /3.0/ Third body: HE
/0.7/ LOW/3.165E23 -1.8E0 0.0E0/ TROE/3.7E-1
3.315E3 6.1E1 1.0E99/
99 CH4+H=CH3+H2 6.14E+05 2.5 9587 REV/6.73E2 2.946E0 8.047E3/
100 CH4+OH=CH3+H2O 5.83E+04 2.6 2190 REV/6.776E2 2.94E0 1.554E4/
101 CH4+O=CH3+OH 1.02E+09 1.5 8600 REV/5.804E5 1.927E0 5.648E3/
102 CH4+HO2=CH3+H2O2 1.13E+01 3.74 21010 REV/7.166E0 3.491E0 3.468E3/
103 CH4+CH2=2CH3 2.46E+06 2 8270 REV/1.736E6 1.868E0 1.298E4/
104 CH3+OH=CH2O+H2 8.00E+09 0.5 -1755 REV/1.066E12 3.22E-1 6.821E4/
105 CH3+OH=CH2(S)+H2O 4.51E+17 -1.34 1417 REV/1.654E16 -8.55E-1 1.039E3/
106 CH3+OH=CH3O+H 6.94E+07 1.343 11200 REV/1.5E12 5.0E-1 -1.1E2/
107 CH3+OH=CH2OH+H 3.09E+07 1.596 4506 REV/1.65E11 6.5E-1 -2.84E2/
108 CH3+OH=CH2+H2O 5.60E+07 1.6 5420 REV/9.224E5 2.072E0 1.406E4/
109 CH3+HO2=CH3O+OH 1.00E+12 0.269 -687.5 REV/6.19E12 1.47E-1 2.455E4/
110 CH3+HO2=CH4+O2 1.16E+05 2.23 -3022 REV/2.018E7 2.132E0 5.321E4/
111 CH3+O=CH2O+H 5.54E+13 0.05 -136 REV/3.83E15 -1.47E-1 6.841E4/
112 CH3+O2=CH3O+O 7.55E+12 0 28320 REV/4.718E14 -4.51E-1 2.88E2/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
113 CH3+O2=CH2O+OH 4.11E+11 0 13840 REV/8.224E10 1.94E-1 6.566E4/
114 CH3+O2(+M)=CH3O2(+M) 1.01E+08 1.63 0 LOW/3.816E31 -4.89E0 3.432E3/ TROE/4.5E-2
8.801E2 2.5E9 1.786E9/
115 CH3O2+CH2O=CH3O2H+HCO 1.99E+12 0 11660 REV/1.323E14 -8.53E-1 9.259E3/
116 CH4+CH3O2=CH3+CH3O2H 1.81E+11 0 18480 REV/2.233E12 -6.94E-1 -6.55E2/
117 CH3OH+CH3O2=CH2OH+CH3O2H 1.81E+12 0 13710 REV/2.346E14 -1.031E0 2.404E3/
118 CH3O2+CH3=2CH3O 5.08E+12 0 -1411 REV/1.967E12 1.76E-1 2.807E4/
119 CH3O2+HO2=CH3O2H+O2 2.47E+11 0 -1570 REV/5.302E14 -7.92E-1 3.552E4/
120 2CH3O2=¿CH2O+CH3OH+O2 3.11E+14 -1.61 -1051
121 2CH3O2=¿O2+2CH3O 1.40E+16 -1.61 1860
122 CH3O2+H=CH3O+OH 9.60E+13 0 0 REV/1.72E9 1.019E0 4.078E4/
123 CH3O2+O=CH3O+O2 3.60E+13 0 0 REV/2.229E11 6.28E-1 5.752E4/
124 CH3O2+OH=CH3OH+O2 6.00E+13 0 0 REV/1.536E13 4.34E-1 5.916E4/
125 CH3O2H=CH3O+OH 6.31E+14 0 42300 REV/2.514E6 1.883E0 -2.875E3/
126 CH2(S)=CH2 1.00E+13 0 0 REV/4.488E12 -1.3E-2 9.02E3/
127 CH2(S)+CH4=2CH3 1.60E+13 0 -570 REV/5.067E12 -1.45E-1 1.316E4/
128 CH2(S)+O2=¿CO+OH+H 7.00E+13 0 0
129 CH2(S)+H2=CH3+H 7.00E+13 0 0 REV/2.022E16 -5.91E-1 1.527E4/
130 CH2(S)+H=CH2+H 3.00E+13 0 0 REV/1.346E13 -1.3E-2 9.02E3/
131 CH2(S)+H=CH+H2 3.00E+13 0 0 REV/6.948E13 -2.53E-1 1.248E4/
132 CH2(S)+O=¿CO+2H 3.00E+13 0 0
133 CH2(S)+OH=CH2O+H 3.00E+13 0 0 REV/1.154E18 -7.7E-1 8.523E4/
134 CH2(S)+CO2=CH2O+CO 3.00E+12 0 0 REV/4.366E10 4.21E-1 5.981E4/
135 CH2+H(+M)=CH3(+M) 2.50E+16 -0.8 0 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.7/
LOW/3.2E27 -3.14E0 1.23E3/ TROE/6.8E-1 7.8E1
1.995E3 5.59E3/
136 CH2+O2=CH2O+O 2.40E+12 0 1500 REV/5.955E14 -3.65E-1 6.098E4/
137 CH2+O2=¿CO2+2H 5.80E+12 0 1500
138 CH2+O2=¿CO+OH+H 5.00E+12 0 1500
139 CH2+O=¿CO+2H 5.00E+13 0 0
140 CH2+H=CH+H2 1.00E+18 -1.56 0 REV/5.16E18 -1.8E0 3.46E3/ DUP
141 CH2+OH=CH+H2O 1.13E+07 2 3000 REV/6.183E8 1.655E0 2.135E4/
142 CH+O2=HCO+O 3.30E+13 0 0 REV/9.371E12 1.61E-1 7.121E4/
249
Reaction A B E Auxiliary data

250
143 C+OH=CO+H 5.00E+13 0 0 REV/1.356E15 0.0E0 1.543E5/
144 C+O2=CO+O 5.00E+13 0 0 REV/1.07E14 0.0E0 1.382E5/
145 CH+H=C+H2 5.00E+13 0 0 REV/2.043E14 0.0E0 2.382E4/
146 CH+O=CO+H 5.70E+13 0 0 REV/2.774E15 0.0E0 1.76E5/
147 CH+OH=HCO+H 3.00E+13 0 0 REV/5.069E14 0.0E0 8.811E4/
148 CH2+H=CH+H2 2.70E+11 0.67 25700 REV/1.897E11 6.7E-1 2.873E4/ DUP
149 CH+H2O=H+CH2O 1.71E+13 0 -755 REV/8.372E14 0.0E0 5.752E4/
150 CH+CO2=HCO+CO 1.70E+12 0 685 REV/2.565E11 0.0E0 6.646E4/
151 2CH3(+M)=C2H6(+M) 9.21E+16 -1.17 635.8 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.7/
LOW/1.135E36 -5.246E0 1.705E3/ TROE/4.05E-1
1.12E3 6.96E1 1.0E10/
152 C2H5+H(+M)=C2H6(+M) 5.21E+17 -0.99 1580 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.7/
LOW/1.99E41 -7.08E0 6.685E3/ TROE/8.42E-1
1.25E2 2.219E3 6.882E3/
153 C2H6+H=C2H5+H2 1.15E+08 1.9 7530 REV/1.062E4 2.582E0 9.76E3/
154 C2H6+O=C2H5+OH 3.55E+06 2.4 5830 REV/1.702E2 3.063E0 6.648E3/
155 C2H6+OH=C2H5+H2O 1.48E+07 1.9 950 REV/1.45E4 2.476E0 1.807E4/
156 C2H6+O2=C2H5+HO2 6.03E+13 0 51870 REV/2.921E10 3.34E-1 -5.93E2/
157 C2H6+CH3=C2H5+CH4 1.51E-07 6 6047 REV/1.273E-8 6.236E0 9.817E3/
158 C2H6+HO2=C2H5+H2O2 3.46E+01 3.61 16920 REV/1.849E0 3.597E0 3.151E3/
159 C2H6+CH3O2=C2H5+CH3O2H 1.94E+01 3.64 17100 REV/2.017E1 3.182E0 1.734E3/
160 C2H6+CH3O=C2H5+CH3OH 2.41E+11 0 7090 REV/4.779E8 4.69E-1 9.547E3/
161 C2H6+CH=C2H5+CH2 1.10E+14 0 -260 REV/1.969E9 9.21E-1 -1.49E3/
162 CH2(S)+C2H6=CH3+C2H5 1.20E+14 0 0 REV/3.203E12 9.1E-2 1.75E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
163 C2H4+H(+M)=C2H5(+M) 8.10E+11 0.454 1820 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4
/2.0/ Third body: C2H6 /3.0/ Third body: HE
/0.7/ LOW/9.0E41 -7.62E0 6.97E3/ TROE/9.75E-1
2.1E2 9.84E2 4.374E3/
164 H2+CH3O2=H+CH3O2H 1.50E+14 0 26030 REV/1.688E18 -1.14E0 8.434E3/
165 H2+C2H5O2=H+C2H5O2H 1.50E+14 0 26030 REV/1.691E18 -1.14E0 8.438E3/
166 C2H5+C2H3=2C2H4 6.86E+11 0.11 -4300 REV/4.82E14 0.0E0 7.153E4/
167 CH3+C2H5=CH4+C2H4 1.18E+04 2.45 -2921 REV/2.39E6 2.4E0 6.669E4/
168 C2H5+H=2CH3 9.69E+13 0 220 REV/2.029E9 1.028E0 1.051E4/
169 C2H5+H=C2H4+H2 2.00E+12 0 0 REV/4.44E11 3.96E-1 6.807E4/
170 C2H5+O=CH3CHO+H 1.10E+14 0 0 REV/1.033E17 -5.0E-1 7.742E4/
171 C2H5+HO2=C2H5O+OH 1.10E+13 0 0 REV/9.68E15 -7.23E-1 2.765E4/
172 CH3O2+C2H5=CH3O+C2H5O 8.00E+12 0 -1000 REV/4.404E14 -4.25E-1 3.089E4/
173 C2H5O+O2=CH3CHO+HO2 4.28E+10 0 1097 REV/1.322E8 6.15E-1 3.413E4/
174 C2H5O=CH3+CH2O 1.32E+20 -2.018 20750 REV/3.0E11 0.0E0 6.336E3/
175 C2H5O=CH3CHO+H 5.43E+15 -0.687 22230 REV/8.0E12 0.0E0 6.4E3/
176 C2H5O2=C2H5+O2 1.31E+62 -14.784 49180 REV/2.876E56 -1.382E1 1.462E4/
177 C2H5O2+CH2O=C2H5O2H+HCO 1.99E+12 0 11660 REV/1.325E14 -8.53E-1 9.263E3/
178 CH4+C2H5O2=CH3+C2H5O2H 1.81E+11 0 18480 REV/2.237E12 -6.94E-1 -6.51E2/
179 CH3OH+C2H5O2=CH2OH+C2H5O2H 1.81E+12 0 13710 REV/2.35E14 -1.031E0 2.408E3/
180 C2H5O2+HO2=C2H5O2H+O2 1.75E+10 0 -3275 REV/3.763E13 -7.92E-1 3.382E4/
181 C2H6+C2H5O2=C2H5+C2H5O2H 8.60E+00 3.76 17200 REV/8.957E0 3.302E0 1.838E3/
182 C2H5O2H=C2H5O+OH 6.31E+14 0 42300 REV/5.661E8 1.033E0 -1.705E3/
183 C2H4O2H=C2H5+O2 4.37E+47 -12.115 31020 REV/1.814E45 -1.15E1 1.46E4/
184 C2H5+O2=C2H4+HO2 7.56E+14 -1.01 4749 REV/8.802E14 -9.62E-1 1.813E4/ DUP
185 C2H5+O2=C2H4+HO2 4.00E-01 3.88 13620 REV/4.656E-1 3.928E0 2.7E4/ DUP
186 C2H5+O2=C2H4O1-2+OH 1.63E+11 -0.31 6150 REV/3.633E13 -6.26E-1 3.984E4/
187 C2H5+O2=CH3CHO+OH 8.27E+02 2.41 5285 REV/2.247E3 2.301E0 6.597E4/
188 C2H5O2=C2H4O2H 2.28E+39 -8.479 45170 REV/1.203E36 -8.13E0 2.702E4/
189 C2H5O2=CH3CHO+OH 2.52E+41 -10.2 43710 REV/1.502E36 -9.345E0 6.984E4/
190 C2H5O2=C2H4+HO2 1.82E+38 -8.45 37890 REV/4.632E32 -7.438E0 1.67E4/
191 C2H5O2=C2H4O1-2+OH 4.00E+43 -10.46 45580 REV/1.959E40 -9.812E0 4.471E4/
192 C2H4O2H=C2H4O1-2+OH 8.85E+30 -6.08 20660 REV/8.199E30 -5.781E0 3.793E4/
193 C2H4O2H=C2H4+HO2 3.98E+34 -7.25 23250 REV/1.922E32 -6.587E0 2.021E4/
194 C2H4O2H=CH3CHO+OH 1.19E+34 -9.02 29210 REV/1.339E32 -8.514E0 7.348E4/
251
Reaction A B E Auxiliary data

252
195 C2H4O1-2=CH3+HCO 3.63E+13 0 57200 REV/1.006E4 1.549E0 -2.75E3/
196 C2H4O1-2=CH3CHO 7.41E+12 0 53800 REV/9.013E10 2.07E-1 8.08E4/
197 C2H4O1-2+OH=C2H3O1-2+H2O 1.78E+13 0 3610 REV/1.347E10 6.93E-1 2.474E4/
198 C2H4O1-2+H=C2H3O1-2+H2 8.00E+13 0 9680 REV/5.71E9 7.99E-1 1.592E4/
199 C2H4O1-2+HO2=C2H3O1-2+H2O2 1.13E+13 0 30430 REV/4.666E11 1.04E-1 2.067E4/
200 C2H4O1-2+CH3O2=C2H3O1-2+CH3O2H 1.13E+13 0 30430 REV/9.078E12 -3.41E-1 1.907E4/
201 C2H4O1-2+C2H5O2=C2H3O1-2+C2H5O2H 1.13E+13 0 30430 REV/9.093E12 -3.41E-1 1.908E4/
202 C2H4O1-2+CH3=C2H3O1-2+CH4 1.07E+12 0 11830 REV/6.967E10 3.53E-1 1.961E4/
203 C2H4O1-2+CH3O=C2H3O1-2+CH3OH 1.20E+11 0 6750 REV/1.839E8 5.86E-1 1.322E4/
204 C2H3O1-2=CH3CO 8.50E+14 0 14000 REV/1.002E14 4.1E-2 4.871E4/
205 C2H3O1-2=CH2CHO 1.00E+14 0 14000 REV/1.245E15 -3.75E-1 4.401E4/
206 CH3CHO=CH3+HCO 7.69E+20 -1.342 86950 REV/1.75E13 0.0E0 0.0E0/
207 CH3CHO+H=CH3CO+H2 1.11E+13 0 3110 REV/7.678E9 6.33E-1 1.706E4/
208 CH3CHO+O=CH3CO+OH 5.94E+12 0 1868 REV/2.133E9 6.14E-1 1.441E4/
209 CH3CHO+OH=CH3CO+H2O 2.00E+06 1.8 1300 REV/1.467E4 2.327E0 3.014E4/
210 CH3CHO+O2=CH3CO+HO2 3.01E+13 0 39150 REV/1.092E11 2.85E-1 -1.588E3/
211 CH3CHO+CH3=CH3CO+CH4 1.76E+03 2.79 4950 REV/1.111E3 2.977E0 2.044E4/
212 CH3CHO+HO2=CH3CO+H2O2 3.01E+12 0 11920 REV/1.205E12 -6.2E-2 9.877E3/
213 CH3O2+CH3CHO=CH3O2H+CH3CO 3.01E+12 0 11920 REV/2.344E13 -5.07E-1 8.282E3/
214 CH3CHO+CH3CO3=CH3CO+CH3CO3H 3.01E+12 0 11920 REV/1.922E12 -1.0E-2 1.265E4/
215 CH3CHO+OH=CH3+HOCHO 3.00E+15 -1.076 0 REV/2.371E16 -1.277E0 2.375E4/
216 CH3CHO+OH=CH2CHO+H2O 1.72E+05 2.4 815 REV/1.332E5 2.511E0 2.495E4/
217 CH3CO(+M)=CH3+CO(+M) 3.00E+12 0 16720 LOW/1.2E15 0.0E0 1.2518E4/
218 CH3CO+H=CH2CO+H2 2.00E+13 0 0 REV/1.037E13 2.01E-1 6.056E4/
219 CH3CO+O=CH2CO+OH 2.00E+13 0 0 REV/5.381E12 1.82E-1 5.914E4/
220 CH3CO+CH3=CH2CO+CH4 5.00E+13 0 0 REV/2.364E16 -2.45E-1 6.21E4/
221 CH3CO3=CH3CO+O2 6.86E+19 -1.949 38530 REV/1.2E11 0.0E0 -1.1E3/
222 CH3CO3+HO2=CH3CO3H+O2 1.75E+10 0 -3275 REV/3.08E12 -2.94E-1 3.818E4/
223 H2O2+CH3CO3=HO2+CH3CO3H 2.41E+12 0 9936 REV/3.845E12 5.3E-2 1.271E4/
224 CH4+CH3CO3=CH3+CH3CO3H 1.81E+11 0 18480 REV/1.831E11 -1.96E-1 3.711E3/
225 CH2O+CH3CO3=HCO+CH3CO3H 1.99E+12 0 11660 REV/1.085E13 -3.56E-1 1.362E4/
226 C2H6+CH3CO3=C2H5+CH3CO3H 1.70E+13 0 20460 REV/1.45E12 4.0E-2 9.46E3/
227 CH3CO3H=CH3CO2+OH 5.01E+14 0 40150 REV/3.618E7 1.761E0 1.338E3/
228 CH3CO2+M=CH3+CO2+M 4.40E+15 0 10500 REV/4.548E8 1.378E0 1.752E4/
229 CH2CHO=CH2CO+H 4.07E+15 -0.342 50600 REV/5.0E13 0.0E0 1.23E4/
230 CH2CHO+O2=CH2CO+HO2 1.81E+11 0 1840 REV/7.151E10 0.0E0 3.385E3/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
231 CH2+CO(+M)=CH2CO(+M) 8.10E+11 0 0 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.7/
LOW/2.69E33 -5.11E0 7.095E3/ TROE/5.907E-1
2.75E2 1.226E3 5.185E3/
232 CH2CO+H=CH3+CO 1.10E+13 0 3400 REV/2.4E12 0.0E0 4.02E4/
233 CH2CO+H=HCCO+H2 2.00E+14 0 8000 REV/1.434E11 4.7E-1 4.52E3/
234 CH2CO+O=CH2+CO2 1.75E+12 0 1350 REV/2.854E9 8.09E-1 4.944E4/
235 CH2CO+O=HCCO+OH 1.00E+13 0 8000 REV/3.723E9 4.52E-1 3.108E3/
236 CH2CO+OH=HCCO+H2O 1.00E+13 0 2000 REV/7.604E10 3.65E-1 1.341E4/
237 CH2CO+OH=CH2OH+CO 2.00E+12 0 -1010 REV/8.17E9 4.94E-1 2.453E4/
238 CH2(S)+CH2CO=C2H4+CO 1.60E+14 0 0 REV/3.75E14 2.17E-1 1.034E5/
239 HCCO+OH=¿H2+2CO 1.00E+14 0 0
240 H+HCCO=CH2(S)+CO 1.10E+13 0 0 REV/4.061E7 1.561E0 1.854E4/
241 HCCO+O=¿H+2CO 8.00E+13 0 0
242 HCCO+O2=¿OH+2CO 4.20E+10 0 850
243 HCCO+M=CH+CO+M 6.50E+15 0 58820 REV/1.391E11 1.033E0 -1.372E4/
244 CH+CH2O=H+CH2CO 9.46E+13 0 -515 REV/1.623E15 0.0E0 6.906E4/
245 CH+HCCO=CO+C2H2 5.00E+13 0 0 REV/1.721E17 0.0E0 1.646E5/
246 C2H3+H(+M)=C2H4(+M) 1.36E+14 0.173 660 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4
/2.0/ Third body: C2H6 /3.0/ Third body: HE
/0.7/ LOW/1.4E30 -3.86E0 3.32E3/ TROE/7.82E-1
2.075E2 2.663E3 6.095E3/
253
Reaction A B E Auxiliary data

254
247 C2H4(+M)=C2H2+H2(+M) 8.00E+12 0.44 88770 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4
/2.0/ Third body: C2H6 /3.0/ Third body: HE
/0.7/ LOW/1.58E51 -9.3E0 9.78E4/ TROE/7.35E-1
1.8E2 1.035E3 5.417E3/
248 C2H4+H=C2H3+H2 5.07E+07 1.93 12950 REV/1.602E4 2.436E0 5.19E3/
249 C2H4+O=CH3+HCO 8.56E+06 1.88 183 REV/3.297E2 2.602E0 2.614E4/
250 C2H4+O=CH2CHO+H 4.99E+06 1.88 183 REV/1.541E9 1.201E0 1.878E4/
251 C2H4+OH=C2H3+H2O 2.09E+06 2.01 1160 REV/7.0E3 2.41E0 8.292E3/
252 C2H4+CH3=C2H3+CH4 6.62E+00 3.7 9500 REV/1.44E0 4.02E0 5.472E3/
253 C2H4+O2=C2H3+HO2 4.00E+13 0 58200 REV/6.626E10 1.58E-1 -4.249E3/
254 C2H4+CH3O=C2H3+CH3OH 1.20E+11 0 6750 REV/8.138E8 2.93E-1 -7.83E2/
255 C2H4+CH3O2=C2H3+CH3O2H 2.23E+12 0 17190 REV/7.929E12 -6.34E-1 -8.167E3/
256 C2H4+C2H5O2=C2H3+C2H5O2H 2.23E+12 0 17190 REV/7.943E12 -6.34E-1 -8.163E3/
257 C2H4+CH3CO3=C2H3+CH3CO3H 1.13E+13 0 30430 REV/3.295E12 -1.36E-1 9.44E3/
258 C2H4+CH3O2=C2H4O1-2+CH3O 2.82E+12 0 17110 REV/3.385E13 -6.5E-2 4.166E4/
259 C2H4+C2H5O2=C2H4O1-2+C2H5O 2.82E+12 0 17110 REV/7.638E15 -9.16E-1 4.283E4/
260 C2H4+HO2=C2H4O1-2+OH 2.23E+12 0 17190 REV/4.28E14 -3.64E-1 3.75E4/
261 CH+CH4=C2H4+H 6.00E+13 0 0 REV/3.573E14 0.0E0 5.548E4/
262 CH2(S)+CH3=C2H4+H 2.00E+13 0 0 REV/6.128E19 -1.223E0 7.305E4/
263 C2H2+H(+M)=C2H3(+M) 5.60E+12 0 2400 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4
/2.0/ Third body: C2H6 /3.0/ Third body: HE
/0.7/ LOW/3.8E40 -7.27E0 7.22E3/ TROE/7.51E-1
9.85E1 1.302E3 4.167E3/
264 C2H3+O2=C2H2+HO2 2.12E-06 6 9484 REV/1.087E-5 5.905E0 2.403E4/
265 C2H3+O2=CH2O+HCO 1.70E+29 -5.312 6500 REV/7.987E27 -4.883E0 9.345E4/
266 C2H3+O2=CH2CHO+O 5.50E+14 -0.611 5260 REV/3.0E18 -1.386E0 1.63E4/
267 CH3+C2H3=CH4+C2H2 3.92E+11 0 0 REV/3.497E14 -1.93E-1 7.078E4/
268 C2H3+H=C2H2+H2 3.00E+13 0 0 REV/2.934E13 2.53E-1 6.924E4/
269 C2H3+OH=C2H2+H2O 5.00E+12 0 0 REV/5.184E13 1.47E-1 8.413E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
270 C2H+H(+M)=C2H2(+M) 1.00E+17 0 0 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4
/2.0/ Third body: C2H6 /3.0/ Third body: HE
/0.7/ LOW/3.75E33 -4.8E0 1.9E3/ TROE/6.46E-1
1.32E2 1.315E3 5.566E3/
271 C2H2+O2=HCCO+OH 2.00E+08 1.5 30100 REV/2.039E6 1.541E0 3.227E4/
272 O+C2H2=C2H+OH 4.60E+19 -1.4 28950 REV/3.023E15 -6.04E-1 -1.782E3/
273 C2H2+O=CH2+CO 6.94E+06 2 1900 REV/4.05E1 3.198E0 4.836E4/
274 C2H2+O=HCCO+H 1.35E+07 2 1900 REV/4.755E7 1.65E0 2.08E4/
275 C2H2+OH=C2H+H2O 3.37E+07 2 14000 REV/4.524E4 2.709E0 -4.28E2/
276 C2H2+OH=CH2CO+H 3.24E+13 0 12000 REV/3.061E17 -8.02E-1 3.579E4/
277 C2H2+OH=CH3+CO 4.83E-04 4 -2000 REV/3.495E-6 4.638E0 5.212E4/
278 OH+C2H2=H+HCCOH 5.04E+05 2.3 13500 REV/3.852E9 1.436E0 4.382E3/
279 H+HCCOH=H+CH2CO 1.00E+13 0 0 REV/9.299E13 -2.9E-1 3.111E4/
280 C2H5OH(+M)=CH2OH+CH3(+M) 5.71E+23 -1.68 94400 Third body: H2 /2.0/ Third body: H2O /5.0/
Third body: CO /2.0/ Third body: CO2 /3.0/
LOW/3.11E85 -1.884E1 1.131E5/ TROE/5.0E-1
5.5E2 8.25E2 6.1E3/
281 C2H5OH(+M)=C2H5+OH(+M) 2.40E+23 -1.62 99540 Third body: H2 /2.0/ Third body: H2O /5.0/
Third body: CO /2.0/ Third body: CO2 /3.0/
LOW/5.11E85 -1.88E1 1.1877E5/ TROE/5.0E-1
6.5E2 8.0E2 1.0E15/
282 C2H5OH(+M)=C2H4+H2O(+M) 2.79E+13 0.09 66140 Third body: H2O /5.0/ LOW/2.57E83 -1.885E1
8.6453E4/ TROE/7.0E-1 3.5E2 8.0E2 3.8E3/
283 C2H5OH(+M)=CH3CHO+H2(+M) 7.24E+11 0.095 91010 Third body: H2O /5.0/ LOW/4.46E87 -1.942E1
1.1558E5/ TROE/9.0E-1 9.0E2 1.1E3 3.5E3/
284 C2H5OH+O2=PC2H4OH+HO2 2.00E+13 0 52800 REV/2.192E10 2.78E-1 4.43E2/
285 C2H5OH+O2=SC2H4OH+HO2 1.50E+13 0 50150 REV/1.946E11 8.9E-2 4.879E3/
286 C2H5OH+OH=PC2H4OH+H2O 1.74E+11 0.27 600 REV/3.857E8 7.9E-1 1.783E4/
287 C2H5OH+OH=SC2H4OH+H2O 4.64E+11 0.15 0 REV/1.217E10 4.81E-1 2.431E4/
288 C2H5OH+OH=C2H5O+H2O 7.46E+11 0.3 1634 REV/3.641E11 4.06E-1 1.631E4/
255
Reaction A B E Auxiliary data

256
289 C2H5OH+H=PC2H4OH+H2 1.23E+07 1.8 5098 REV/2.571E3 2.426E0 7.432E3/
290 C2H5OH+H=SC2H4OH+H2 2.58E+07 1.65 2827 REV/6.383E4 2.087E0 1.225E4/
291 C2H5OH+H=C2H5O+H2 1.50E+07 1.6 3038 REV/6.904E5 1.812E0 2.821E3/
292 C2H5OH+HO2=PC2H4OH+H2O2 1.23E+04 2.55 15750 REV/1.488E3 2.481E0 2.083E3/
293 C2H5OH+HO2=SC2H4OH+H2O2 8.20E+03 2.55 10750 REV/1.174E4 2.292E0 4.169E3/
294 C2H5OH+HO2=C2H5O+H2O2 2.50E+12 0 24000 REV/6.658E13 -4.83E-1 7.782E3/
295 C2H5OH+CH3O2=PC2H4OH+CH3O2H 1.23E+04 2.55 15750 REV/2.894E4 2.036E0 4.88E2/
296 C2H5OH+CH3O2=SC2H4OH+CH3O2H 8.20E+03 2.55 10750 REV/2.284E5 1.847E0 2.574E3/
297 C2H5OH+CH3O2=C2H5O+CH3O2H 2.50E+12 0 24000 REV/1.295E15 -9.27E-1 6.187E3/
298 C2H5OH+O=PC2H4OH+OH 9.41E+07 1.7 5459 REV/1.021E4 2.307E0 6.381E3/
299 C2H5OH+O=SC2H4OH+OH 1.88E+07 1.85 1824 REV/2.415E4 2.268E0 9.832E3/
300 C2H5OH+O=C2H5O+OH 1.58E+07 2 4448 REV/3.775E5 2.194E0 2.819E3/
301 C2H5OH+CH3=PC2H4OH+CH4 1.33E+02 3.18 9362 REV/2.537E1 3.36E0 1.324E4/
302 C2H5OH+CH3=SC2H4OH+CH4 4.44E+02 2.9 7690 REV/1.002E3 2.891E0 1.865E4/
303 C2H5OH+CH3=C2H5O+CH4 1.34E+02 2.92 7452 REV/5.627E3 2.686E0 8.775E3/
304 C2H5OH+C2H5=PC2H4OH+C2H6 5.00E+10 0 13400 REV/6.995E10 0.0E0 2.699E4/
305 C2H5OH+C2H5=SC2H4OH+C2H6 5.00E+10 0 10400 REV/6.995E10 0.0E0 2.399E0/
306 PC2H4OH=C2H4+OH 1.05E+25 -3.99 30390 REV/4.17E20 -2.84E0 1.24E3/
307 SC2H4OH+M=CH3CHO+H+M 1.00E+14 0 25000 REV/2.742E12 4.62E-1 -4.7E2/
308 O2C2H4OH=PC2H4OH+O2 3.90E+16 -1 30000 REV/1.2E11 0.0E0 -1.1E3/
309 O2C2H4OH=¿OH+2CH2O 3.13E+09 0 18900
310 SC2H4OH+O2=CH3CHO+HO2 3.81E+06 2 1641 REV/2.19E5 2.39E0 2.504E4/
311 CH3COCH3(+M)=CH3CO+CH3(+M) 7.11E+21 -1.57 84680 LOW/7.013E89 -2.038E1 1.0715E5/ TROE/8.63E-1
1.0E10 4.164E2 3.29E9/
312 CH3COCH3+OH=CH3COCH2+H2O 1.25E+05 2.483 445 REV/8.62E4 2.322E0 2.471E4/
313 CH3COCH3+H=CH3COCH2+H2 9.80E+05 2.43 5160 REV/6.374E4 2.375E0 1.453E4/
314 CH3COCH3+O=CH3COCH2+OH 5.13E+11 0.211 4890 REV/1.732E10 1.37E-1 1.285E4/
315 CH3COCH3+CH3=CH3COCH2+CH4 3.96E+11 0 9784 REV/2.35E13 -5.01E-1 2.069E4/
316 CH3COCH3+CH3O=CH3COCH2+CH3OH 4.34E+11 0 6460 REV/6.06E11 -2.68E-1 1.606E4/
317 CH3COCH3+O2=CH3COCH2+HO2 6.03E+13 0 48500 REV/2.057E13 -4.03E-1 3.181E3/
318 CH3COCH3+HO2=CH3COCH2+H2O2 1.70E+13 0 20460 REV/6.397E14 -7.5E-1 1.383E4/
319 CH3COCH3+CH3O2=CH3COCH2+CH3O2H 1.70E+13 0 20460 REV/1.245E16 -1.195E0 1.223E4/
320 CH3COCH2=CH2CO+CH3 1.00E+14 0 31000 REV/1.0E11 0.0E0 6.0E3/
321 CH3COCH2O2=CH3COCH2+O2 2.02E+15 -0.956 24460 REV/1.2E11 0.0E0 -1.1E3/
322 CH3COCH3+CH3COCH2O2=CH3COCH2+CH3COCH2O2H 1.00E+11 0 5000 REV/1.995E10 0.0E0 1.0E4/
323 CH2O+CH3COCH2O2=HCO+CH3COCH2O2H 1.29E+11 0 9000 REV/2.512E10 0.0E0 1.01E4/
324 HO2+CH3COCH2O2=¿CH3COCH2O2H+O2 1.00E+12 0 0
325 CH3COCH2O2H=CH3COCH2O+OH 1.00E+16 0 43000 REV/4.242E8 1.74E0 -4.342E3/
326 CH3COCH2O=CH3CO+CH2O 3.73E+20 -2.176 17260 REV/1.0E11 0.0E0 1.19E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
327 C2H3CHO=C2H3+HCO 2.00E+24 -2.135 103400 REV/1.81E13 0.0E0 0.0E0/
328 C2H3CHO+H=C2H3CO+H2 1.34E+13 0 3300 REV/3.311E10 6.13E-1 2.268E4/
329 C2H3CHO+O=C2H3CO+OH 5.94E+12 0 1868 REV/7.618E9 5.94E-1 1.984E4/
330 C2H3CHO+OH=C2H3CO+H2O 9.24E+06 1.5 -962 REV/2.42E5 2.007E0 3.331E4/
331 C2H3CHO+O2=C2H3CO+HO2 1.01E+13 0 40700 REV/1.302E11 2.65E-1 5.391E3/
332 C2H3CHO+HO2=C2H3CO+H2O2 3.01E+12 0 11920 REV/4.303E12 -8.2E-2 1.53E4/
333 C2H3CHO+CH3=C2H3CO+CH4 2.61E+06 1.78 5911 REV/5.878E6 1.947E0 2.683E4/
334 C2H3CHO+C2H3=C2H3CO+C2H4 1.74E+12 0 8440 REV/1.0E13 0.0E0 2.8E4/
335 C2H3CHO+CH3O=C2H3CO+CH3OH 1.00E+12 0 3300 REV/5.304E10 4.01E-1 2.291E4/
336 C2H3CHO+CH3O2=C2H3CO+CH3O2H 3.01E+12 0 11920 REV/8.371E13 -5.27E-1 1.371E4/
337 C2H3CO=C2H3+CO 1.37E+21 -2.179 39410 REV/1.51E11 0.0E0 4.81E3/
338 C2H5CHO=C2H5+HCO 1.50E+27 -3.205 87040 REV/1.81E13 0.0E0 0.0E0/
339 C2H5CHO+H=C2H5CO+H2 4.00E+13 0 4200 REV/2.377E10 6.54E-1 1.813E4/
340 C2H5CHO+O=C2H5CO+OH 5.00E+12 0 1790 REV/1.542E9 6.36E-1 1.431E4/
341 C2H5CHO+OH=C2H5CO+H2O 2.69E+10 0.76 -340 REV/1.695E8 1.308E0 2.848E4/
342 C2H5CHO+CH3=C2H5CO+CH4 2.61E+06 1.78 5911 REV/1.414E6 1.988E0 2.138E4/
343 C2H5CHO+HO2=C2H5CO+H2O2 2.80E+12 0 13600 REV/9.626E11 -4.1E-2 1.153E4/
344 C2H5CHO+CH3O=C2H5CO+CH3OH 1.00E+12 0 3300 REV/1.276E10 4.42E-1 1.746E4/
345 C2H5CHO+CH3O2=C2H5CO+CH3O2H 3.01E+12 0 11920 REV/2.013E13 -4.85E-1 8.26E3/
346 C2H5CHO+C2H5=C2H5CO+C2H6 1.00E+12 0 8000 REV/6.432E12 -2.8E-2 1.97E4/
347 C2H5CHO+C2H5O=C2H5CO+C2H5OH 6.03E+11 0 3300 REV/3.02E11 0.0E0 1.816E4/
348 C2H5CHO+C2H5O2=C2H5CO+C2H5O2H 3.01E+12 0 11920 REV/2.017E13 -4.86E-1 8.264E3/
349 C2H5CHO+O2=C2H5CO+HO2 1.01E+13 0 40700 REV/3.131E10 3.06E-1 -5.8E1/
350 C2H5CHO+CH3CO3=C2H5CO+CH3CO3H 3.01E+12 0 11920 REV/1.651E12 1.2E-2 1.263E4/
351 C2H5CHO+C2H3=C2H5CO+C2H4 1.70E+12 0 8440 REV/3.198E12 1.48E-1 3.013E4/
352 C2H5CO=C2H5+CO 2.46E+23 -3.208 17550 REV/1.51E11 0.0E0 4.81E3/
353 CH3OCH3(+M)=CH3+CH3O(+M) 7.25E+21 -0.94 80250 LOW/3.5E60 -1.156E1 1.01E5/ TROE/1.83E-1
1.3E0 1.3E4 6.71E9/
354 CH3OCH3+OH=CH3OCH2+H2O 6.32E+06 2 -651.7 REV/7.853E4 2.236E0 2.121E4/
355 CH3OCH3+H=CH3OCH2+H2 7.72E+06 2.09 3384 REV/9.042E3 2.432E0 1.036E4/
356 CH3OCH3+O=CH3OCH2+OH 7.75E+08 1.36 2250 REV/4.712E5 1.683E0 7.81E3/
357 CH3OCH3+HO2=CH3OCH2+H2O2 1.68E+13 0 17690 REV/1.138E13 -3.53E-1 8.657E3/
358 CH3OCH3+CH3O2=CH3OCH2+CH3O2H 1.68E+13 0 17690 REV/2.215E14 -7.98E-1 7.062E3/
359 CH3OCH3+CH3=CH3OCH2+CH4 1.45E-06 5.73 5700 REV/1.544E-6 5.626E0 1.421E4/
360 CH3OCH3+O2=CH3OCH2+HO2 4.10E+13 0 44910 REV/2.518E11 -6.0E-3 -2.806E3/
361 CH3OCH3+CH3O=CH3OCH2+CH3OH 6.02E+11 0 4074 REV/7.383E10 -2.7E-1 1.026E4/
362 CH3OCH3+CH3OCH2O2=CH3OCH2+CH3OCH2O2H 5.00E+12 0 17690 REV/6.428E13 -7.94E-1 7.258E3/
363 CH3OCH3+O2CHO=CH3OCH2+HO2CHO 4.43E+04 2.6 13910 REV/1.651E-5 4.607E0 1.172E3/
364 CH3OCH3+OCHO=CH3OCH2+HOCHO 1.00E+13 0 17690 REV/7.195E12 -3.14E-1 3.313E4/
257
Reaction A B E Auxiliary data

258
365 CH3OCH2=CH2O+CH3 1.60E+13 0 25500 REV/2.601E5 1.879E0 1.667E4/
366 CH3OCH2+CH3O=CH3OCH3+CH2O 2.41E+13 0 0 REV/1.25E14 3.2E-1 7.854E4/
367 CH3OCH2+CH2O=CH3OCH3+HCO 5.49E+03 2.8 5862 REV/2.768E4 2.745E0 1.408E4/
368 CH3OCH2+CH3CHO=CH3OCH3+CH3CO 1.26E+12 0 8499 REV/7.746E11 2.8E-1 1.698E4/
369 CH3OCH2O2=CH3OCH2+O2 4.44E+19 -1.594 36240 REV/2.0E12 0.0E0 0.0E0/
370 CH3OCH2O2+CH2O=CH3OCH2O2H+HCO 1.00E+12 0 11660 REV/6.482E13 -8.49E-1 9.455E3/
371 CH3OCH2O2+CH3CHO=CH3OCH2O2H+CH3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
372 2CH3OCH2O2=¿O2+2CH3OCH2O 2.21E+23 -4.5 0
373 CH3OCH2O2H=CH3OCH2O+OH 2.11E+22 -2.124 43830 REV/2.0E13 0.0E0 0.0E0/
374 CH3OCH2O=CH3O+CH2O 4.38E+19 -2.014 25190 REV/1.0E11 0.0E0 1.19E4/
375 CH3OCH2O+O2=CH3OCHO+HO2 5.00E+10 0 500 REV/1.086E10 -2.0E-2 4.648E4/
376 CH3OCH2O=CH3OCHO+H 6.06E+12 0.056 8218 REV/1.0E13 0.0E0 7.838E3/
377 CH3OCH2O2=CH2OCH2O2H 6.00E+10 0 21580 REV/1.248E12 -7.65E-1 1.112E4/
378 CH2OCH2O2H=¿OH+2CH2O 1.50E+13 0 20760
379 O2CH2OCH2O2H=CH2OCH2O2H+O2 1.92E+19 -1.622 36270 REV/7.0E11 0.0E0 0.0E0/
380 O2CH2OCH2O2H=HO2CH2OCHO+OH 4.00E+10 0 18580 REV/5.041E3 1.416E0 5.964E4/
381 HO2CH2OCHO=OCH2OCHO+OH 2.00E+16 0 40500 REV/1.183E8 1.934E0 -3.952E3/
382 OCH2OCHO=CH2O+OCHO 2.90E+19 -2.201 31850 REV/1.25E11 0.0E0 1.19E4/
383 OCH2OCHO=HOCH2OCO 1.00E+11 0 14000 REV/1.568E9 4.87E-1 2.067E4/
384 HOCH2OCO=HOCH2O+CO 2.24E+19 -2.021 19690 REV/1.5E11 0.0E0 4.8E3/
385 HOCH2OCO=CH2OH+CO2 2.41E+17 -1.574 22120 REV/1.5E11 0.0E0 3.572E4/
386 CH3OCHO=CH2OCHO+H 8.24E+19 -1.15 102500 REV/1.0E14 0.0E0 0.0E0/
387 CH3OCHO=CH3OCO+H 1.33E+19 -1 100100 REV/1.0E14 0.0E0 0.0E0/
388 CH3OCHO(+M)=CH3OH+CO(+M) 1.00E+14 0 62500 LOW/6.143E60 -1.207E1 7.54E4/ TROE/7.8E-1
8.28E9 4.389E2 6.7E8/
389 CH3OCHO=CH3O+HCO 5.37E+16 -0.01 97090 REV/3.0E13 0.0E0 0.0E0/
390 CH3OCHO=CH3+OCHO 3.21E+17 -0.53 79970 REV/1.0E13 0.0E0 0.0E0/
391 CH3OCHO+O2=CH3OCO+HO2 1.00E+13 0 49700 REV/2.979E10 3.29E-1 9.998E-1/
392 CH3OCHO+O2=CH2OCHO+HO2 2.05E+13 0 52000 REV/2.143E10 3.56E-1 3.31E2/
393 CH3OCHO+OH=CH3OCO+H2O 1.58E+07 1.8 934 REV/9.519E4 2.371E0 2.082E4/
394 CH3OCHO+OH=CH2OCHO+H2O 5.27E+09 0.97 1586 REV/1.114E7 1.568E0 1.95E4/
395 CH3OCHO+HO2=CH3OCO+H2O2 4.82E+03 2.6 13910 REV/1.585E3 2.581E0 2.899E3/
396 CH3OCHO+HO2=CH2OCHO+H2O2 2.38E+04 2.55 16490 REV/2.746E3 2.558E0 3.513E3/
397 CH3OCHO+O=CH3OCO+OH 2.76E+05 2.45 2830 REV/8.126E1 3.108E0 6.408E3/
398 CH3OCHO+O=CH2OCHO+OH 9.80E+05 2.43 4750 REV/1.014E2 3.115E0 6.358E3/
399 CH3OCHO+H=CH3OCO+H2 6.50E+05 2.4 4471 REV/3.693E2 3.076E0 9.461E3/
400 CH3OCHO+H=CH2OCHO+H2 6.65E+05 2.54 6756 REV/1.326E2 3.243E0 9.776E3/
401 CH3OCHO+CH3=CH3OCO+CH4 7.55E-01 3.46 5481 REV/3.914E-1 3.691E0 1.201E4/
402 CH3OCHO+CH3=CH2OCHO+CH4 4.52E-01 3.65 7154 REV/8.222E-2 3.908E0 1.171E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
403 CH3OCHO+CH3O=CH3OCO+CH3OH 5.48E+11 0 5000 REV/6.685E9 4.64E-1 1.022E4/
404 CH3OCHO+CH3O=CH2OCHO+CH3OH 2.17E+11 0 6458 REV/9.289E8 4.91E-1 9.705E3/
405 CH3OCHO+CH3O2=CH3OCO+CH3O2H 4.82E+03 2.6 13910 REV/3.083E4 2.137E0 1.304E3/
406 CH3OCHO+CH3O2=CH2OCHO+CH3O2H 2.38E+04 2.55 16490 REV/5.342E4 2.114E0 1.918E3/
407 CH3OCHO+HCO=CH3OCO+CH2O 5.40E+06 1.9 17010 REV/5.196E5 2.29E0 6.806E3/
408 CH3OCHO+HCO=CH2OCHO+CH2O 1.03E+05 2.5 18430 REV/3.461E3 2.917E0 6.261E3/
409 CH2OCHO=CH3OCO 2.62E+11 -0.03 38180 REV/1.629E12 -1.8E-1 4.067E4/
410 CH3OCO=CH3+CO2 3.59E+14 -0.172 16010 REV/4.76E7 1.54E0 3.47E4/
411 CH3OCO=CH3O+CO 1.43E+15 -0.041 23770 REV/1.55E6 2.02E0 5.73E3/
412 CH2OCHO=CH2O+HCO 4.66E+12 0.12 27440 REV/1.5E11 0.0E0 1.19E4/
413 C3H8(+M)=CH3+C2H5(+M) 1.29E+37 -5.84 97380 Third body: H2 /2.0/ Third body: H2O /6.0/
Third body: AR /0.7/ Third body: CO /1.5/
Third body: CO2 /2.0/ Third body: CH4 /2.0/
Third body: C2H6 /3.0/ Third body: HE /0.7/
LOW/5.64E74 -1.574E1 9.8714E4/ TROE/3.1E-1
5.0E1 3.0E3 9.0E3/
414 C3H8=NC3H7+H 3.75E+17 -0.357 101200 REV/1.0E14 0.0E0 0.0E0/
415 C3H8=IC3H7+H 2.38E+18 -0.671 98680 REV/1.0E14 0.0E0 0.0E0/
416 C3H8+O2=IC3H7+HO2 2.00E+13 0 49640 REV/1.764E9 5.99E-1 -1.69E2/
417 C3H8+O2=NC3H7+HO2 6.00E+13 0 52290 REV/3.354E10 2.85E-1 -5.9E1/
418 H+C3H8=H2+IC3H7 1.30E+06 2.4 4471 REV/2.186E1 3.347E0 9.351E3/
419 H+C3H8=H2+NC3H7 1.33E+06 2.54 6756 REV/1.418E2 3.173E0 9.096E3/
420 C3H8+O=IC3H7+OH 5.49E+05 2.5 3140 REV/4.793E0 3.428E0 6.608E3/
421 C3H8+O=NC3H7+OH 3.71E+06 2.4 5505 REV/2.053E2 3.014E0 6.433E3/
422 C3H8+OH=NC3H7+H2O 1.05E+10 0.97 1586 REV/1.191E7 1.497E0 1.882E4/
423 C3H8+OH=IC3H7+H2O 4.67E+07 1.61 -35 REV/8.327E3 2.451E0 1.974E4/
424 C3H8+HO2=IC3H7+H2O2 5.88E+04 2.5 14860 REV/5.721E2 2.752E0 3.742E3/
425 C3H8+HO2=NC3H7+H2O2 8.10E+04 2.5 16690 REV/4.995E3 2.438E0 3.03E3/
426 CH3+C3H8=CH4+IC3H7 6.40E+04 2.17 7520 REV/9.819E2 2.671E0 1.394E4/
427 CH3+C3H8=CH4+NC3H7 9.04E-01 3.65 7154 REV/8.791E-2 3.837E0 1.103E4/
428 IC3H7+C3H8=NC3H7+C3H8 3.00E+10 0 12900 REV/3.0E10 0.0E0 1.29E4/
429 C2H3+C3H8=C2H4+IC3H7 1.00E+11 0 10400 REV/1.31E11 0.0E0 1.78E4/
430 C2H3+C3H8=C2H4+NC3H7 1.00E+11 0 10400 REV/1.31E11 0.0E0 1.78E4/
431 C2H5+C3H8=C2H6+IC3H7 1.00E+11 0 10400 REV/3.63E10 0.0E0 9.934E3/
432 C2H5+C3H8=C2H6+NC3H7 1.00E+11 0 10400 REV/3.63E10 0.0E0 9.934E3/
433 C3H8+C3H5-A=NC3H7+C3H6 7.94E+11 0 20500 REV/5.372E16 -1.33E0 1.34E4/
434 C3H8+C3H5-A=IC3H7+C3H6 7.94E+11 0 16200 REV/5.372E16 -1.33E0 9.095E3/
259
Reaction A B E Auxiliary data

260
435 C3H8+CH3O=NC3H7+CH3OH 3.00E+11 0 7000 REV/1.22E10 0.0E0 9.182E3/
436 C3H8+CH3O=IC3H7+CH3OH 3.00E+11 0 7000 REV/1.22E10 0.0E0 9.182E3/
437 CH3O2+C3H8=CH3O2H+NC3H7 8.10E+04 2.5 16690 REV/9.718E4 1.993E0 1.435E3/
438 CH3O2+C3H8=CH3O2H+IC3H7 5.88E+04 2.5 14860 REV/1.113E4 2.307E0 2.147E3/
439 C2H5O2+C3H8=C2H5O2H+NC3H7 8.10E+04 2.5 16690 REV/9.735E4 1.993E0 1.439E3/
440 C2H5O2+C3H8=C2H5O2H+IC3H7 5.88E+04 2.5 14860 REV/1.115E4 2.307E0 2.151E3/
441 NC3H7O2+C3H8=NC3H7O2H+NC3H7 1.70E+13 0 20460 REV/2.086E13 -5.1E-1 5.0E3/
442 NC3H7O2+C3H8=NC3H7O2H+IC3H7 2.00E+12 0 17000 REV/3.871E11 -1.96E-1 4.08E3/
443 IC3H7O2+C3H8=IC3H7O2H+NC3H7 1.70E+13 0 20460 REV/2.093E13 -5.11E-1 5.0E3/
444 IC3H7O2+C3H8=IC3H7O2H+IC3H7 2.00E+12 0 17000 REV/3.885E11 -1.97E-1 4.08E3/
445 C3H8+CH3CO3=IC3H7+CH3CO3H 2.00E+12 0 17000 REV/3.104E10 3.05E-1 8.65E3/
446 C3H8+CH3CO3=NC3H7+CH3CO3H 1.70E+13 0 20460 REV/1.673E12 -9.0E-3 9.57E3/
447 C3H8+O2CHO=NC3H7+HO2CHO 5.52E+04 2.55 16480 REV/1.187E-8 5.54E0 -1.92E3/
448 C3H8+O2CHO=IC3H7+HO2CHO 1.48E+04 2.6 13910 REV/7.838E-6 4.65E0 -3.0E1/
449 IC3H7=H+C3H6 6.92E+13 -0.025 37690 REV/2.64E13 0.0E0 2.16E3/
450 IC3H7+H=C2H5+CH3 2.00E+13 0 0 REV/4.344E7 1.176E0 8.62E3/
451 IC3H7+O2=C3H6+HO2 4.50E-19 0 5020 REV/2.0E-19 0.0E0 1.75E4/
452 IC3H7+OH=C3H6+H2O 2.41E+13 0 0 REV/2.985E12 5.7E-1 8.382E4/
453 IC3H7+O=CH3COCH3+H 4.82E+13 0 0 REV/1.293E16 -1.9E-1 7.938E4/
454 IC3H7+O=CH3CHO+CH3 4.82E+13 0 0 REV/1.279E11 8.0E-1 8.648E4/
455 NC3H7=CH3+C2H4 9.97E+40 -8.6 41430 REV/1.898E34 -6.99E0 1.71E4/
456 NC3H7=H+C3H6 8.78E+39 -8.1 46580 REV/2.07E37 -7.39E0 1.202E4/
457 NC3H7+O2=C3H6+HO2 3.00E-19 0 3000 REV/2.0E-19 0.0E0 1.75E4/
458 C2H5CHO+NC3H7=C2H5CO+C3H8 1.70E+12 0 8440 REV/1.9E14 0.0E0 1.879E4/
459 C2H5CHO+IC3H7=C2H5CO+C3H8 1.70E+12 0 8440 REV/1.9E14 0.0E0 1.879E4/
460 C2H5CHO+C3H5-A=C2H5CO+C3H6 1.70E+12 0 8440 REV/1.0E13 0.0E0 2.8E4/
461 C3H6=C2H3+CH3 2.73E+62 -13.28 123200 REV/6.822E53 -1.1779E1 2.055E4/
462 C3H6=C3H5-A+H 2.01E+61 -13.26 118500 REV/2.041E61 -1.352E1 3.061E4/
463 C3H6=C3H5-S+H 7.71E+69 -16.09 140000 REV/2.551E67 -1.5867E1 2.869E4/
464 C3H6=C3H5-T+H 5.62E+71 -16.58 139300 REV/4.26E68 -1.6164E1 3.008E4/
465 C3H6+O=C2H5+HCO 1.58E+07 1.76 -1216 REV/9.188E1 2.725E0 2.311E4/
466 C3H6+O=¿CH2CO+CH3+H 2.50E+07 1.76 76
467 C3H6+O=¿CH3CHCO+2H 2.50E+07 1.76 76
468 C3H6+O=C3H5-A+OH 5.24E+11 0.7 5884 REV/1.104E11 6.97E-1 2.015E4/
469 C3H6+O=C3H5-S+OH 1.20E+11 0.7 8959 REV/8.239E7 1.18E0 -2.07E2/
470 C3H6+O=C3H5-T+OH 6.03E+10 0.7 7632 REV/9.483E6 1.373E0 5.76E2/
471 C3H6+OH=C3H5-A+H2O 3.12E+06 2 -298 REV/1.343E7 1.909E0 3.027E4/
472 C3H6+OH=C3H5-S+H2O 2.11E+06 2 2778 REV/2.959E4 2.393E0 9.916E3/
473 C3H6+OH=C3H5-T+H2O 1.11E+06 2 1451 REV/3.565E3 2.586E0 1.07E4/
474 C3H6+HO2=C3H5-A+H2O2 2.70E+04 2.5 12340 REV/6.341E6 1.82E0 1.201E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
475 C3H6+HO2=C3H5-S+H2O2 1.80E+04 2.5 27620 REV/1.377E4 2.304E0 3.864E3/
476 C3H6+HO2=C3H5-T+H2O2 9.00E+03 2.5 23590 REV/1.577E3 2.497E0 1.941E3/
477 C3H6+H=C3H5-A+H2 1.73E+05 2.5 2492 REV/7.023E4 2.515E0 1.817E4/
478 C3H6+H=C3H5-S+H2 8.04E+05 2.5 12280 REV/1.063E3 2.999E0 4.526E3/
479 C3H6+H=C3H5-T+H2 4.05E+05 2.5 9794 REV/1.227E2 3.192E0 4.15E3/
480 C3H6+H=C2H4+CH3 2.30E+13 0 2547 REV/7.272E7 1.271E0 1.12E4/
481 C3H6+O2=C3H5-A+HO2 4.00E+12 0 39900 REV/8.514E12 -3.33E-1 8.87E2/
482 C3H6+O2=C3H5-S+HO2 2.00E+12 0 62900 REV/1.387E10 1.51E-1 4.59E2/
483 C3H6+O2=C3H5-T+HO2 1.40E+12 0 60700 REV/2.224E9 3.44E-1 3.69E2/
484 C3H6+CH3=C3H5-A+CH4 2.21E+00 3.5 5675 REV/8.184E2 3.07E0 2.289E4/
485 C3H6+CH3=C3H5-S+CH4 1.35E+00 3.5 12850 REV/1.626E0 3.553E0 6.635E3/
486 C3H6+CH3=C3H5-T+CH4 8.40E-01 3.5 11660 REV/2.322E-1 3.746E0 7.552E3/
487 C3H6+C2H5=C3H5-A+C2H6 1.00E+11 0 9800 REV/5.369E5 1.33E0 1.644E4/
488 C3H6+CH3CO3=C3H5-A+CH3CO3H 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
489 C3H6+CH3O2=C3H5-A+CH3O2H 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
490 C3H6+HO2=C3H6O1-2+OH 1.29E+12 0 14900 REV/1.0E-10 0.0E0 0.0E0/
491 C3H6+C2H5O2=C3H5-A+C2H5O2H 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
492 C3H6+NC3H7O2=C3H5-A+NC3H7O2H 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
493 C3H6+IC3H7O2=C3H5-A+IC3H7O2H 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
494 C3H6OH=C3H6+OH 4.34E+15 -0.805 27900 REV/9.93E11 0.0E0 -9.6E2/
495 HOC3H6O2=C3H6OH+O2 2.87E+19 -1.897 34290 REV/1.2E11 0.0E0 -1.1E3/
496 HOC3H6O2=¿CH3CHO+CH2O+OH 1.25E+10 0 18900
497 C3H5-A=C2H2+CH3 2.40E+48 -9.9 82080 REV/2.61E46 -9.82E0 3.695E4/
498 C3H5-A=C3H4-A+H 4.19E+13 0.216 61930 REV/2.4E11 6.9E-1 3.007E3/
499 C3H5-A+HO2=C3H5O+OH 7.00E+12 0 -1000 REV/1.605E12 6.0E-2 1.166E4/
500 C3H5-A+CH3O2=C3H5O+CH3O 7.00E+12 0 -1000 REV/1.99E15 -7.4E-1 1.702E4/
501 C3H5-A+H=C3H4-A+H2 1.23E+03 3.035 2582 REV/2.818E0 3.784E0 4.722E4/
502 C3H5-A+CH3=C3H4-A+CH4 1.00E+11 0 0 REV/4.921E12 5.0E-2 4.778E4/
503 C3H5-A+C2H5=C2H6+C3H4-A 4.00E+11 0 0 REV/1.802E12 5.0E-2 4.033E4/
504 C3H5-A+C2H5=C2H4+C3H6 4.00E+11 0 0 REV/6.937E16 -1.33E0 5.28E4/
505 C3H5-A+C2H3=C2H4+C3H4-A 1.00E+12 0 0 REV/1.624E13 5.0E-2 4.819E4/
506 C3H4-A+C3H6=2C3H5-A 8.39E+17 -1.29 33690 REV/1.0E12 0.0E0 0.0E0/
507 C3H5-A+O2=C3H4-A+HO2 2.18E+21 -2.85 30760 REV/2.614E19 -2.449E0 2.071E4/
508 C3H5-A+O2=CH2CHO+CH2O 7.14E+15 -1.21 21050 REV/4.944E16 -1.4E0 8.862E4/
509 C3H5-A+O2=C2H3CHO+OH 2.47E+13 -0.44 23020 REV/1.989E13 -6.09E-1 7.514E4/
510 C3H5-A+O2=¿C2H2+CH2O+OH 9.72E+29 -5.71 21450
511 C3H5-S=C2H2+CH3 9.60E+39 -8.17 42030 REV/1.61E40 -8.58E0 2.033E4/
512 C3H5-S=C3H4-P+H 4.19E+15 -0.79 37480 REV/5.8E12 0.0E0 3.1E3/
513 C3H5-S+O2=CH3CHO+HCO 4.34E+12 0 0 REV/1.611E17 -1.27E0 9.653E4/
514 C3H5-S+H=C3H4-A+H2 3.33E+12 0 0 REV/7.977E12 1.1E-1 6.886E4/
261
Reaction A B E Auxiliary data

262
515 C3H5-S+CH3=C3H4-A+CH4 1.00E+11 0 0 REV/6.253E12 1.1E-1 6.934E4/
516 C3H5-T=C2H2+CH3 2.16E+40 -8.31 45110 REV/1.61E40 -8.58E0 2.033E4/
517 C3H5-T=C3H4-A+H 3.51E+14 -0.44 40890 REV/8.5E12 0.0E0 2.0E3/
518 C3H5-T=C3H4-P+H 1.08E+15 -0.6 38490 REV/6.5E12 0.0E0 2.0E3/
519 C3H5-T+O2=C3H4-A+HO2 1.89E+30 -5.59 15540 REV/3.037E31 -5.865E0 2.681E4/
520 C3H5-T+O2=CH3COCH2+O 3.81E+17 -1.36 5580 REV/2.0E11 0.0E0 1.75E4/
521 C3H5-T+O2=CH2O+CH3CO 3.71E+25 -3.96 7043 REV/1.872E27 -4.43E0 1.012E5/
522 C3H5-T+H=C3H4-P+H2 3.33E+12 0 0 REV/2.138E16 -8.8E-1 7.105E4/
523 C3H5-T+CH3=C3H4-P+CH4 1.00E+11 0 0 REV/1.676E16 -8.8E-1 7.153E4/
524 C3H4-A+M=C3H3+H+M 1.14E+17 0 70000 REV/1.798E15 -3.8E-1 1.061E4/
525 C3H4-A=C3H4-P 1.20E+15 0 92400 REV/3.222E18 -9.9E-1 9.659E4/
526 C3H4-A+O2=C3H3+HO2 4.00E+13 0 39160 REV/3.17E11 -8.6E-2 3.11E2/
527 C3H4-A+HO2=CH2CO+CH2+OH 4.00E+12 0 19000 REV/1.0E0 0.0E0 0.0E0/
528 C3H4-A+OH=CH2CO+CH3 3.12E+12 0 -397 REV/1.806E17 -1.38E0 3.607E4/
529 C3H4-A+OH=C3H3+H2O 1.00E+07 2 1000 REV/1.602E5 2.157E0 3.173E4/
530 C3H4-A+O=C2H4+CO 7.80E+12 0 1600 REV/3.269E8 1.252E0 1.219E5/
531 C3H4-A+O=C2H2+CH2O 3.00E-03 4.61 -4243 REV/2.32E2 3.23E0 8.119E4/
532 C3H4-A+H=C3H3+H2 2.00E+07 2 5000 REV/3.022E4 2.262E0 2.084E4/
533 C3H4-A+CH3=C3H3+CH4 3.67E-02 4.01 6830 REV/5.06E-2 3.826E0 2.421E4/
534 C3H4-A+C3H5-A=C3H3+C3H6 2.00E+11 0 7700 REV/2.644E19 -2.71E0 4.214E4/
535 C3H4-A+C2H=C3H3+C2H2 1.00E+13 0 0 REV/1.42E16 -1.38E0 5.382E4/
536 C3H4-P+M=C3H3+H+M 1.14E+17 0 70000 REV/6.708E11 6.1E-1 6.42E3/
537 C3H4-P=C2H+CH3 4.20E+16 0 100000 REV/1.018E12 6.1E-1 -1.6E3/
538 C3H4-P+O2=HCCO+OH+CH2 1.00E+07 1.5 30100 REV/1.0E0 0.0E0 0.0E0/
539 C3H4-P+O2=C3H3+HO2 2.00E+13 0 41600 REV/6.371E11 -2.08E-1 1.021E3/
540 C3H4-P+HO2=C2H4+CO+OH 3.00E+12 0 19000 REV/1.0E0 0.0E0 0.0E0/
541 C3H4-P+OH=C3H3+H2O 1.00E+07 2 1000 REV/6.441E5 2.034E0 3.0E4/
542 C3H4-P+OH=CH2CO+CH3 5.00E-04 4.5 -1000 REV/1.079E-2 4.11E0 3.128E4/
543 C3H4-P+O=C2H3+HCO 3.20E+12 0 2010 REV/2.548E12 -3.9E-1 3.235E4/
544 C3H4-P+O=HCCO+CH3 9.60E+08 1 0 REV/1.43E4 1.793E0 2.699E4/
545 C3H4-P+O=HCCO+CH2+H 3.20E-19 0 2010 REV/1.0E-30 0.0E0 0.0E0/
546 C3H4-P+O=C3H3+OH 7.65E+08 1.5 8600 REV/2.177E8 1.31E0 2.247E4/
547 C3H4-P+H=C3H3+H2 2.00E+07 2 5000 REV/1.215E5 2.14E0 1.911E4/
548 C3H4-P+CH3=C3H3+CH4 1.50E+00 3.5 5600 REV/8.313E0 3.195E0 2.125E4/
549 C3H4-P+C2H=C3H3+C2H2 1.00E+12 0 0 REV/5.297E11 -3.9E-1 4.963E4/
550 C3H4-P+C2H3=C3H3+C2H4 1.00E+12 0 7700 REV/9.541E11 -3.9E-1 5.245E4/
551 C3H4-P+C3H5-A=C3H3+C3H6 1.00E+12 0 7700 REV/4.931E16 -1.73E0 3.795E4/
552 C3H3+O=CH2O+C2H 1.00E+13 0 0 REV/5.446E14 0.0E0 3.161E4/
553 C3H3+OH=C3H2+H2O 1.00E+13 0 0 REV/1.343E15 0.0E0 1.568E4/
554 C3H3+O2=CH2CO+HCO 3.01E+10 0 2870 REV/4.881E11 0.0E0 5.947E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
555 C3H3+CH3=C2H5+C2H 4.30E+15 -0.79 45630 REV/1.81E13 0.0E0 0.0E0/
556 C3H2+O2=HCO+HCCO 5.00E+13 0 0 REV/2.326E14 -2.14E-1 7.719E4/
557 C3H4-A+HO2=C2H4+CO+OH 1.00E+12 0 14000 REV/1.0E0 0.0E0 0.0E0/
558 C3H4-A+HO2=C3H3+H2O2 3.00E+13 0 14000 REV/1.551E16 -1.38E0 4.4E4/
559 C2H2+CH3=C3H4-P+H 4.23E+08 1.143 12090 REV/1.0E14 0.0E0 4.0E3/
560 C2H2+CH3=C3H4-A+H 6.74E+19 -2.08 31590 REV/6.407E25 -3.345E0 2.177E4/
561 C3H3+H=C3H2+H2 5.00E+13 0 0 REV/5.999E7 1.365E0 4.11E3/
562 C3H2+OH=C2H2+HCO 5.00E+13 0 0 REV/2.282E16 -2.54E-1 7.502E4/
563 C3H2+O2=¿HCCO+CO+H 5.00E+13 0 0
564 CH3CHCO+OH=¿C2H5+CO2 1.73E+12 0 -1010
565 CH3CHCO+OH=¿SC2H4OH+CO 2.00E+12 0 -1010
566 CH3CHCO+H=¿C2H5+CO 4.40E+12 0 1459
567 CH3CHCO+O=¿CH3CHO+CO 3.20E+12 0 -437
568 NC3H7+HO2=NC3H7O+OH 7.00E+12 0 -1000 REV/6.22E15 -6.92E-1 2.531E4/
569 IC3H7+HO2=IC3H7O+OH 7.00E+12 0 -1000 REV/1.051E16 -5.57E-1 2.732E4/
570 CH3O2+NC3H7=CH3O+NC3H7O 7.00E+12 0 -1000 REV/3.89E14 -3.94E-1 2.955E4/
571 CH3O2+IC3H7=CH3O+IC3H7O 7.00E+12 0 -1000 REV/6.573E14 -2.58E-1 3.156E4/
572 NC3H7O2=NC3H7+O2 2.40E+20 -1.616 35960 REV/4.52E12 0.0E0 0.0E0/
573 IC3H7O2=IC3H7+O2 3.13E+22 -2.167 38160 REV/7.54E12 0.0E0 0.0E0/
574 NC3H7O2+CH2O=NC3H7O2H+HCO 5.60E+12 0 13600 REV/8.0E11 0.0E0 1.0E4/
575 NC3H7O2+CH3CHO=NC3H7O2H+CH3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
576 IC3H7O2+CH2O=IC3H7O2H+HCO 5.60E+12 0 13600 REV/8.0E11 0.0E0 1.0E4/
577 IC3H7O2+CH3CHO=IC3H7O2H+CH3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
578 NC3H7O2+HO2=NC3H7O2H+O2 1.75E+10 0 -3275 REV/3.841E13 -7.95E-1 3.361E4/
579 IC3H7O2+HO2=IC3H7O2H+O2 1.75E+10 0 -3275 REV/3.855E13 -7.96E-1 3.361E4/
580 C2H4+NC3H7O2=C2H3+NC3H7O2H 1.13E+13 0 30430 REV/3.0E12 0.0E0 1.15E4/
581 C2H4+IC3H7O2=C2H3+IC3H7O2H 1.13E+13 0 30430 REV/3.0E12 0.0E0 1.15E4/
582 CH3OH+NC3H7O2=CH2OH+NC3H7O2H 6.30E+12 0 19360 REV/1.0E9 0.0E0 1.0E4/
583 CH3OH+IC3H7O2=CH2OH+IC3H7O2H 6.30E+12 0 19360 REV/1.0E9 0.0E0 1.0E4/
584 C2H3CHO+NC3H7O2=C2H3CO+NC3H7O2H 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
585 C2H3CHO+IC3H7O2=C2H3CO+IC3H7O2H 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
586 CH4+NC3H7O2=CH3+NC3H7O2H 1.12E+13 0 24640 REV/7.43E11 0.0E0 5.5E3/
587 CH4+IC3H7O2=CH3+IC3H7O2H 1.12E+13 0 24640 REV/7.43E11 0.0E0 5.5E3/
588 NC3H7O2+CH3O2=¿NC3H7O+CH3O+O2 1.40E+16 -1.61 1860
589 IC3H7O2+CH3O2=¿IC3H7O+CH3O+O2 1.40E+16 -1.61 1860
590 H2+NC3H7O2=H+NC3H7O2H 3.01E+13 0 26030 REV/4.8E13 0.0E0 7.95E3/
591 H2+IC3H7O2=H+IC3H7O2H 3.01E+13 0 26030 REV/4.8E13 0.0E0 7.95E3/
592 IC3H7O2+C2H6=IC3H7O2H+C2H5 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
593 NC3H7O2+C2H6=NC3H7O2H+C2H5 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
594 IC3H7O2+C2H5CHO=IC3H7O2H+C2H5CO 2.00E+11 0 9500 REV/5.0E9 0.0E0 1.0E4/
263
Reaction A B E Auxiliary data

264
595 NC3H7O2+C2H5CHO=NC3H7O2H+C2H5CO 2.00E+11 0 9500 REV/5.0E9 0.0E0 1.0E4/
596 IC3H7O2+CH3CO3=¿IC3H7O+CH3CO2+O2 1.40E+16 -1.61 1860
597 NC3H7O2+CH3CO3=¿NC3H7O+CH3CO2+O2 1.40E+16 -1.61 1860
598 IC3H7O2+C2H5O2=¿IC3H7O+C2H5O+O2 1.40E+16 -1.61 1860
599 NC3H7O2+C2H5O2=¿NC3H7O+C2H5O+O2 1.40E+16 -1.61 1860
600 2IC3H7O2=¿O2+2IC3H7O 1.40E+16 -1.61 1860
601 2NC3H7O2=¿O2+2NC3H7O 1.40E+16 -1.61 1860
602 IC3H7O2+NC3H7O2=¿IC3H7O+NC3H7O+O2 1.40E+16 -1.61 1860
603 IC3H7O2+CH3=IC3H7O+CH3O 7.00E+12 0 -1000 REV/1.145E11 6.95E-1 2.672E4/
604 IC3H7O2+C2H5=IC3H7O+C2H5O 7.00E+12 0 -1000 REV/1.628E13 9.4E-2 2.913E4/
605 IC3H7O2+IC3H7=2IC3H7O 7.00E+12 0 -1000 REV/2.778E13 2.61E-1 2.98E4/
606 IC3H7O2+NC3H7=IC3H7O+NC3H7O 7.00E+12 0 -1000 REV/1.644E13 1.25E-1 2.78E4/
607 IC3H7O2+C3H5-A=IC3H7O+C3H5O 7.00E+12 0 -1000 REV/4.242E9 8.77E-1 1.414E4/
608 NC3H7O2+CH3=NC3H7O+CH3O 7.00E+12 0 -1000 REV/5.303E12 9.0E-3 2.692E4/
609 NC3H7O2+C2H5=NC3H7O+C2H5O 7.00E+12 0 -1000 REV/7.54E14 -5.92E-1 2.933E4/
610 NC3H7O2+IC3H7=NC3H7O+IC3H7O 7.00E+12 0 -1000 REV/1.286E15 -4.25E-1 3.0E4/
611 NC3H7O2+NC3H7=2NC3H7O 7.00E+12 0 -1000 REV/7.612E14 -5.61E-1 2.8E4/
612 NC3H7O2+C3H5-A=NC3H7O+C3H5O 7.00E+12 0 -1000 REV/1.964E11 1.91E-1 1.434E4/
613 NC3H7O2H=NC3H7O+OH 1.50E+16 0 42500 REV/1.143E8 1.719E0 -4.034E3/
614 IC3H7O2H=IC3H7O+OH 9.45E+15 0 42600 REV/1.55E6 2.406E0 -4.132E3/
615 NC3H7O=C2H5+CH2O 2.72E+21 -2.449 15700 REV/1.0E11 0.0E0 3.496E3/
616 NC3H7O=C2H5CHO+H 8.90E+10 0.746 19800 REV/4.0E12 0.0E0 6.26E3/
617 IC3H7O=CH3+CH3CHO 5.33E+19 -1.696 17140 REV/1.0E11 0.0E0 9.256E3/
618 IC3H7O=CH3COCH3+H 8.66E+14 -0.483 20080 REV/2.0E12 0.0E0 7.27E3/
619 IC3H7O+O2=CH3COCH3+HO2 9.09E+09 0 390 REV/1.0E11 0.0E0 3.2E4/
620 NC3H7O2=C3H6OOH1-2 6.00E+11 0 26850 REV/1.117E8 5.83E-1 1.172E4/
621 NC3H7O2=C3H6OOH1-3 1.13E+11 0 24400 REV/2.716E11 -5.07E-1 8.936E3/
622 IC3H7O2=C3H6OOH2-1 1.80E+12 0 29400 REV/1.122E10 1.19E-1 1.181E4/
623 IC3H7O2=C3H6OOH2-2 1.23E+35 -6.96 48880 REV/2.384E34 -7.06E0 4.494E4/
624 C3H6OOH1-2=C3H6O1-2+OH 6.00E+11 0 22000 REV/1.15E11 4.9E-1 3.837E4/
625 C3H6OOH1-3=C3H6O1-3+OH 7.50E+10 0 15250 REV/1.186E6 1.765E0 2.871E4/
626 C3H6OOH2-1=C3H6O1-2+OH 6.00E+11 0 22000 REV/2.78E8 1.191E0 3.609E4/
627 C3H6OOH1-2=C3H6+HO2 7.83E+15 -1.3 15950 REV/1.0E11 0.0E0 1.1E4/
628 C3H6OOH2-1=C3H6+HO2 3.24E+18 -2 18970 REV/1.0E11 0.0E0 1.175E4/
629 C3H6OOH1-3=¿OH+CH2O+C2H4 3.04E+15 -0.79 27400
630 C3H6OOH2-1=C2H3OOH+CH3 6.54E+27 -5.14 38320 REV/4.46E22 -4.24E0 1.063E4/
631 C3H6OOH1-2=¿C2H4+CH2O+OH 1.31E+33 -7.01 48120
632 C3H6OOH2-2=CH3COCH3+OH 9.00E+14 0 1500 REV/1.021E14 3.1E-1 3.675E4/
633 C3H6OOH1-2O2=C3H6OOH1-2+O2 2.39E+25 -2.945 40100 REV/5.0E12 0.0E0 0.0E0/
634 C3H6OOH1-3O2=C3H6OOH1-3+O2 2.85E+20 -1.626 35690 REV/4.52E12 0.0E0 0.0E0/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
635 C3H6OOH2-1O2=C3H6OOH2-1+O2 5.23E+22 -2.244 37820 REV/4.52E12 0.0E0 0.0E0/
636 C3H6OOH1-2O2=C3KET12+OH 6.00E+11 0 26400 REV/9.249E4 1.329E0 4.892E4/
637 C3H6OOH1-3O2=C3KET13+OH 7.50E+10 0 21400 REV/4.101E3 1.496E0 4.474E4/
638 C3H6OOH2-1O2=C3KET21+OH 3.00E+11 0 23850 REV/1.397E3 1.834E0 4.975E4/
639 C3H6OOH2-1O2=C3H51-2,3OOH 1.13E+11 0 24400 REV/2.391E11 -4.99E-1 8.92E3/
640 C3H6OOH1-2O2=C3H51-2,3OOH 9.00E+11 0 29400 REV/1.913E12 -4.99E-1 1.392E4/
641 C3H51-2,3OOH=AC3H5OOH+HO2 2.56E+13 -0.49 17770 REV/3.18E15 -1.16E0 1.204E4/
642 C3H52-1,3OOH=C3H6OOH1-3O2 1.26E+12 -0.36 13940 REV/6.0E11 0.0E0 2.685E4/
643 C3H52-1,3OOH=AC3H5OOH+HO2 1.15E+14 -0.63 17250 REV/1.564E11 1.2E-1 1.02E4/
644 C3KET12=¿CH3CHO+HCO+OH 9.45E+15 0 43000
645 C3KET13=¿CH2O+CH2CHO+OH 1.00E+16 0 43000
646 C3KET21=¿CH2O+CH3CO+OH 1.00E+16 0 43000
647 AC3H5OOH=C3H5O+OH 3.88E+19 -1.46 45370 REV/2.0E13 0.0E0 0.0E0/
648 C3H5O=C2H3CHO+H 1.00E+14 0 29100 REV/1.676E14 -1.56E-1 1.969E4/
649 C3H5O=C2H3+CH2O 1.46E+20 -1.968 35090 REV/1.5E11 0.0E0 1.06E4/
650 C3H5O+O2=C2H3CHO+HO2 1.00E+12 0 6000 REV/1.288E11 0.0E0 3.2E4/
651 C2H3OOH=CH2CHO+OH 8.40E+14 0 43000 REV/1.0E11 0.0E0 0.0E0/
652 C3H6O1-2=C2H4+CH2O 6.00E+14 0 60000 REV/2.97E11 1.0E0 3.108E4/
653 C3H6O1-2+OH=¿CH2O+C2H3+H2O 5.00E+12 0 0
654 C3H6O1-2+H=¿CH2O+C2H3+H2 2.63E+07 2 5000
655 C3H6O1-2+O=¿CH2O+C2H3+OH 8.43E+13 0 5200
656 C3H6O1-2+HO2=¿CH2O+C2H3+H2O2 1.00E+13 0 15000
657 C3H6O1-2+CH3O2=¿CH2O+C2H3+CH3O2H 1.00E+13 0 19000
658 C3H6O1-2+CH3=¿CH2O+C2H3+CH4 2.00E+11 0 10000
659 C3H6O1-3=C2H4+CH2O 6.00E+14 0 60000 REV/2.97E11 0.0E0 3.108E4/
660 C3H6O1-3+OH=¿CH2O+C2H3+H2O 5.00E+12 0 0
661 C3H6O1-3+O=¿CH2O+C2H3+OH 8.43E+13 0 5200
662 C3H6O1-3+H=¿CH2O+C2H3+H2 2.63E+07 2 5000
663 C3H6O1-3+CH3O2=¿CH2O+C2H3+CH3O2H 1.00E+13 0 19000
664 C3H6O1-3+HO2=¿CH2O+C2H3+H2O2 1.00E+13 0 15000
665 C3H6O1-3+CH3=¿CH2O+C2H3+CH4 2.00E+11 0 10000
666 IC3H7O2=C3H6+HO2 1.02E+43 -9.409 41490 REV/1.954E33 -7.289E0 1.667E4/
667 NC3H7O2=C3H6+HO2 5.04E+38 -8.112 40490 REV/1.198E30 -6.229E0 2.042E4/
668 C4H10(+M)=2C2H5(+M) 2.72E+15 0 75610 LOW/4.72E18 0.0E0 4.9576E4/ TROE/7.2E-1
1.5E3 1.0E-10 1.0E10/
669 C4H10(+M)=NC3H7+CH3(+M) 4.28E+14 0 69900 LOW/5.34E17 0.0E0 4.2959E4/ TROE/7.2E-1
1.5E3 1.0E-10 1.0E10/
670 C4H10=PC4H9+H 1.34E+17 -0.356 101200 REV/3.61E13 0.0E0 0.0E0/
265

671 C4H10=SC4H9+H 1.98E+18 -0.694 98720 REV/3.61E13 0.0E0 0.0E0/


Reaction A B E Auxiliary data

266
672 C4H10+O2=PC4H9+HO2 6.00E+13 0 52340 REV/3.377E10 2.84E-1 -1.9E1/
673 C4H10+O2=SC4H9+HO2 4.00E+13 0 49800 REV/1.532E9 6.22E-1 -4.9E1/
674 C4H10+C3H5-A=PC4H9+C3H6 7.94E+11 0 20500 REV/1.0E12 0.0E0 2.0E4/
675 C4H10+C3H5-A=SC4H9+C3H6 3.16E+11 0 16400 REV/1.0E12 0.0E0 2.0E4/
676 C4H10+C2H5=PC4H9+C2H6 1.58E+11 0 12300 REV/3.56E10 0.0E0 1.292E4/
677 C4H10+C2H5=SC4H9+C2H6 1.00E+11 0 10400 REV/7.12E10 0.0E0 9.917E3/
678 C4H10+C2H3=PC4H9+C2H4 1.00E+12 0 18000 REV/2.57E12 0.0E0 2.538E4/
679 C4H10+C2H3=SC4H9+C2H4 8.00E+11 0 16800 REV/2.05E12 0.0E0 2.418E4/
680 C4H10+CH3=PC4H9+CH4 9.04E-01 3.65 7154 REV/8.853E-2 3.836E0 1.102E4/
681 C4H10+CH3=SC4H9+CH4 3.02E+00 3.46 5481 REV/2.013E-2 3.984E0 1.186E4/
682 C4H10+H=PC4H9+H2 1.88E+05 2.75 6280 REV/2.018E1 3.382E0 8.61E3/
683 C4H10+H=SC4H9+H2 2.60E+06 2.4 4471 REV/1.9E1 3.37E0 9.311E3/
684 C4H10+OH=PC4H9+H2O 1.05E+10 0.97 1586 REV/1.2E7 1.496E0 1.881E4/
685 C4H10+OH=SC4H9+H2O 9.34E+07 1.61 -35 REV/7.235E3 2.474E0 1.97E4/
686 C4H10+O=PC4H9+OH 1.13E+14 0 7850 REV/1.48E13 0.0E0 1.224E4/
687 C4H10+O=SC4H9+OH 5.62E+13 0 5200 REV/7.35E12 0.0E0 9.59E3/
688 C4H10+HO2=PC4H9+H2O2 8.10E+04 2.5 16690 REV/5.031E3 2.437E0 3.02E3/
689 C4H10+HO2=SC4H9+H2O2 1.18E+05 2.5 14860 REV/4.971E2 2.775E0 3.702E3/
690 C4H10+CH3O=PC4H9+CH3OH 3.00E+11 0 7000 REV/1.22E10 0.0E0 5.0E4/
691 C4H10+CH3O=SC4H9+CH3OH 6.00E+11 0 7000 REV/2.441E10 0.0E0 5.0E4/
692 C4H10+C2H5O=PC4H9+C2H5OH 3.00E+11 0 7000 REV/1.0E10 0.0E0 9.0E3/
693 C4H10+C2H5O=SC4H9+C2H5OH 6.00E+11 0 7000 REV/1.0E10 0.0E0 9.0E3/
694 C4H10+PC4H9=SC4H9+C4H10 1.00E+11 0 10400 REV/1.5E11 0.0E0 1.23E4/
695 C4H10+CH3CO3=PC4H9+CH3CO3H 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
696 C4H10+CH3CO3=SC4H9+CH3CO3H 1.12E+13 0 17700 REV/5.0E11 0.0E0 6.5E3/
697 C4H10+O2CHO=PC4H9+HO2CHO 1.68E+13 0 20440 REV/3.68E0 2.99E0 2.505E3/
698 C4H10+O2CHO=SC4H9+HO2CHO 1.12E+13 0 17690 REV/7.595E3 2.06E0 4.266E3/
699 CH3O2+C4H10=CH3O2H+PC4H9 8.10E+04 2.5 16690 REV/9.787E4 1.992E0 1.425E3/
700 CH3O2+C4H10=CH3O2H+SC4H9 1.18E+05 2.5 14860 REV/9.671E3 2.33E0 2.107E3/
701 C2H5O2+C4H10=C2H5O2H+PC4H9 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
702 C2H5O2+C4H10=C2H5O2H+SC4H9 1.12E+13 0 17700 REV/5.0E11 0.0E0 6.5E3/
703 NC3H7O2+C4H10=NC3H7O2H+PC4H9 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
704 NC3H7O2+C4H10=NC3H7O2H+SC4H9 1.12E+13 0 17700 REV/5.0E11 0.0E0 6.5E3/
705 IC3H7O2+C4H10=IC3H7O2H+PC4H9 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
706 IC3H7O2+C4H10=IC3H7O2H+SC4H9 1.12E+13 0 17700 REV/5.0E11 0.0E0 6.5E3/
707 PC4H9O2+C3H8=PC4H9O2H+NC3H7 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
708 PC4H9O2+C3H8=PC4H9O2H+IC3H7 2.00E+12 0 17000 REV/5.0E11 0.0E0 6.5E3/
709 PC4H9O2+C4H10=PC4H9O2H+PC4H9 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
710 PC4H9O2+C4H10=PC4H9O2H+SC4H9 1.12E+13 0 17700 REV/5.0E11 0.0E0 6.5E3/
711 SC4H9O2+C3H8=SC4H9O2H+NC3H7 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
712 SC4H9O2+C3H8=SC4H9O2H+IC3H7 2.00E+12 0 17000 REV/5.0E11 0.0E0 6.5E3/
713 SC4H9O2+C4H10=SC4H9O2H+PC4H9 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
714 SC4H9O2+C4H10=SC4H9O2H+SC4H9 1.12E+13 0 17700 REV/5.0E11 0.0E0 6.5E3/
715 PC4H9=C2H5+C2H4 3.50E+12 0.463 29470 REV/1.32E4 2.48E0 6.13E3/
716 SC4H9=C3H6+CH3 4.80E+10 1.044 30350 REV/1.76E4 2.48E0 6.13E3/
717 PC4H9=C4H8-1+H 2.62E+12 0.253 35700 REV/2.5E11 5.1E-1 2.62E3/
718 SC4H9=C4H8-2+H 2.84E+11 0.337 35520 REV/2.5E11 5.1E-1 2.62E3/
719 SC4H9=C4H8-1+H 3.03E+11 0.591 36820 REV/4.24E11 5.1E-1 1.23E3/
720 PC4H9+O2=C4H8-1+HO2 2.00E-18 0 5000 REV/2.0E-19 0.0E0 1.75E4/
721 SC4H9+O2=C4H8-1+HO2 2.00E-18 0 5000 REV/2.0E-19 0.0E0 1.75E4/
722 SC4H9+O2=C4H8-2+HO2 2.00E-18 0 5000 REV/2.0E-19 0.0E0 1.75E4/
723 C4H8-1=C3H5-A+CH3 1.50E+19 -1 73400 REV/1.35E13 0.0E0 0.0E0/
724 C4H8-1=C2H3+C2H5 1.00E+19 -1 96770 REV/9.0E12 0.0E0 0.0E0/
725 C4H8-1=H+C4H71-3 4.11E+18 -1 97350 REV/5.0E13 0.0E0 0.0E0/
726 C4H8-1+O2=C4H71-3+HO2 2.00E+13 0 37190 REV/4.653E12 7.0E-2 -1.68E2/
727 C4H8-1+H=C4H71-1+H2 7.81E+05 2.5 12290 REV/2.213E5 2.36E0 6.469E3/
728 C4H8-1+H=C4H71-2+H2 3.90E+05 2.5 5821 REV/2.558E4 2.55E0 2.125E3/
729 C4H8-1+H=C4H71-3+H2 3.38E+05 2.36 207 REV/4.323E6 2.1E0 2.033E4/
730 C4H8-1+H=C4H71-4+H2 6.65E+05 2.54 6756 REV/3.045E4 2.54E0 1.103E4/
731 C4H8-1+OH=C4H71-1+H2O 2.14E+06 2 2778 REV/2.625E6 1.86E0 1.212E4/
732 C4H8-1+OH=C4H71-2+H2O 2.22E+06 2 1451 REV/6.304E5 2.05E0 1.291E4/
733 C4H8-1+OH=C4H71-3+H2O 2.76E+04 2.64 -1919 REV/1.532E6 2.38E0 3.336E4/
734 C4H8-1+OH=C4H71-4+H2O 5.27E+09 0.97 1586 REV/1.044E9 9.7E-1 2.101E4/
735 C4H8-1+CH3=C4H71-3+CH4 3.69E+00 3.31 4002 REV/1.234E3 3.05E0 2.461E4/
736 C4H8-1+CH3=C4H71-4+CH4 4.52E-01 3.65 7154 REV/5.405E-1 3.65E0 1.191E4/
737 C4H8-1+HO2=C4H71-3+H2O2 4.82E+03 2.55 10530 REV/1.586E6 1.96E0 1.435E4/
738 C4H8-1+HO2=C4H71-4+H2O2 2.38E+03 2.55 16490 REV/2.8E3 2.22E0 4.46E3/
739 C4H8-1+CH3O2=C4H71-3+CH3O2H 4.82E+03 2.55 10530 REV/3.303E6 1.79E0 1.133E4/
740 C4H8-1+CH3O2=C4H71-4+CH3O2H 2.38E+03 2.55 16490 REV/5.831E3 2.04E0 1.44E3/
741 C4H8-1+CH3O=C4H71-3+CH3OH 4.00E+01 2.9 8609 REV/2.47E2 2.67E0 2.7E4/
742 C4H8-1+CH3O=C4H71-4+CH3OH 2.17E+11 0 6458 REV/4.789E9 2.0E-2 9.002E3/
743 C4H8-1+CH3CO3=C4H71-3+CH3CO3H 1.00E+11 0 8000 REV/2.0E10 0.0E0 1.0E4/
744 C4H8-1+C3H5-A=C4H71-3+C3H6 7.90E+10 0 12400 REV/1.0E11 0.0E0 1.75E4/
745 C4H8-1+C4H6=2C4H71-3 2.35E+12 0 46720 REV/1.6E12 0.0E0 0.0E0/
746 C4H8-1+C2H5O2=C4H71-3+C2H5O2H 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
747 C4H8-1+NC3H7O2=C4H71-3+NC3H7O2H 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
748 C4H8-1+IC3H7O2=C4H71-3+IC3H7O2H 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
749 C4H8-1+PC4H9O2=C4H71-3+PC4H9O2H 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
750 C4H8-1+SC4H9O2=C4H71-3+SC4H9O2H 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
751 C4H8-1+CH3O2=¿C4H8O1-2+CH3O 1.00E+12 0 14340
267
Reaction A B E Auxiliary data

268
752 C4H8-2=H+C4H71-3 4.11E+18 -1 97350 REV/5.0E13 0.0E0 0.0E0/
753 C4H8-2+O2=C4H71-3+HO2 4.00E+13 0 39390 REV/1.35E13 -1.8E-1 -9.24E2/
754 C4H8-2+H=C4H71-3+H2 3.46E+05 2.5 2492 REV/6.428E6 1.99E0 1.966E4/
755 C4H8-2+OH=C4H71-3+H2O 6.24E+06 2 -298 REV/5.019E8 1.49E0 3.202E4/
756 C4H8-2+CH3=C4H71-3+CH4 4.42E+00 3.5 5675 REV/2.145E3 2.99E0 2.332E4/
757 C4H8-2+HO2=C4H71-3+H2O2 1.93E+04 2.6 13910 REV/9.205E6 1.77E0 1.477E4/
758 C4H8-2+CH3O2=C4H71-3+CH3O2H 1.93E+04 2.6 13910 REV/1.917E7 1.59E0 1.175E4/
759 C4H8-2+CH3O=C4H71-3+CH3OH 1.80E+01 2.95 11990 REV/1.612E2 2.47E0 2.742E4/
760 C4H8-2+C2H5O2=C4H71-3+C2H5O2H 3.20E+12 0 14900 REV/1.585E11 0.0E0 1.47E4/
761 C4H8-2+NC3H7O2=C4H71-3+NC3H7O2H 3.20E+12 0 14900 REV/1.585E11 0.0E0 1.47E4/
762 C4H8-2+IC3H7O2=C4H71-3+IC3H7O2H 3.20E+12 0 14900 REV/1.585E11 0.0E0 1.47E4/
763 C4H8-2+PC4H9O2=C4H71-3+PC4H9O2H 3.20E+12 0 14900 REV/1.585E11 0.0E0 1.47E4/
764 C4H8-2+SC4H9O2=C4H71-3+SC4H9O2H 3.20E+12 0 14900 REV/1.585E11 0.0E0 1.47E4/
765 C4H8-1+HO2=¿C4H8O1-2+OH 1.00E+12 0 14340
766 C4H8-2+HO2=¿C4H8O2-3+OH 5.62E+11 0 12310
767 C4H8-2+CH3O2=¿C4H8O2-3+CH3O 5.62E+11 0 12310
768 C4H8OH-1=C4H8-1+OH 1.08E+16 -0.699 28090 REV/4.75E12 0.0E0 -7.82E2/
769 C4H8OH-2=C4H8-2+OH 3.38E+17 -1.253 29920 REV/4.75E12 0.0E0 -7.82E2/
770 C4H8OH-1O2=C4H8OH-1+O2 6.75E+20 -1.944 35520 REV/2.0E12 0.0E0 0.0E0/
771 C4H8OH-2O2=C4H8OH-2+O2 7.69E+20 -1.968 35510 REV/2.0E12 0.0E0 0.0E0/
772 C4H8OH-1O2=¿C2H5CHO+CH2O+OH 1.00E+16 0 25000
773 C4H8OH-2O2=¿OH+2CH3CHO 1.00E+16 0 25000
774 C4H71-1=C2H2+C2H5 1.07E+15 -0.56 30320 REV/2.0E11 0.0E0 7.8E3/
775 C4H71-2=C3H4-A+CH3 9.59E+14 -0.71 31260 REV/2.0E11 0.0E0 7.8E3/
776 C4H71-4=C2H4+C2H3 8.77E+12 -0.22 36290 REV/2.0E11 0.0E0 7.8E3/
777 C4H72-2=C3H4-P+CH3 6.33E+10 0.52 30020 REV/1.0E11 0.0E0 7.8E3/
778 C4H71-3=C4H6+H 1.20E+14 0 49300 REV/4.0E13 0.0E0 1.3E3/
779 C4H71-3+C2H5=C4H8-1+C2H4 2.59E+12 0 -131 REV/1.149E13 6.0E-2 4.944E4/
780 C4H71-3+CH3O=C4H8-1+CH2O 2.41E+13 0 0 REV/2.482E12 2.8E-1 6.633E4/
781 C4H71-3+O=C2H3CHO+CH3 6.03E+13 0 0 REV/3.385E15 -7.8E-1 8.163E4/
782 C4H71-3+HO2=C4H7O+OH 9.64E+12 0 0 REV/7.29E15 -1.09E0 1.553E4/
783 C4H71-3+CH3O2=C4H7O+CH3O 9.64E+12 0 0 REV/7.12E17 -1.67E0 2.029E4/
784 C3H5-A+C4H71-3=C3H6+C4H6 6.31E+12 0 0 REV/1.0E10 0.0E0 5.0E4/
785 C4H71-3+O2=C4H6+HO2 1.00E+09 0 0 REV/1.0E11 0.0E0 1.7E4/
786 H+C4H71-3=C4H6+H2 3.16E+13 0 0 REV/1.066E13 0.0E0 5.681E4/
787 C2H5+C4H71-3=C4H6+C2H6 3.98E+12 0 0 REV/3.211E12 0.0E0 4.984E4/
788 C2H3+C4H71-3=C2H4+C4H6 3.98E+12 0 0 REV/1.157E13 0.0E0 5.771E4/
789 C4H71-3+C2H5O2=C4H7O+C2H5O 3.80E+12 0 -1200 REV/2.0E10 0.0E0 0.0E0/
790 IC3H7O2+C4H71-3=IC3H7O+C4H7O 3.80E+12 0 -1200 REV/2.0E10 0.0E0 0.0E0/
791 NC3H7O2+C4H71-3=NC3H7O+C4H7O 3.80E+12 0 -1200 REV/2.0E10 0.0E0 0.0E0/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
792 C4H7O=CH3CHO+C2H3 7.94E+14 0 19000 REV/1.0E10 0.0E0 2.0E4/
793 C4H7O=C2H3CHO+CH3 7.94E+14 0 19000 REV/1.0E10 0.0E0 2.0E4/
794 C4H6=2C2H3 4.03E+19 -1 98150 REV/1.26E13 0.0E0 0.0E0/
795 C4H6+OH=C2H5+CH2CO 1.00E+12 0 0 REV/3.73E12 0.0E0 3.002E4/
796 C4H6+OH=CH2O+C3H5-A 1.00E+12 0 0 REV/3.501E6 0.0E0 7.106E4/
797 C4H6+OH=C2H3+CH3CHO 1.00E+12 0 0 REV/5.437E11 0.0E0 1.855E4/
798 C4H6+O=C2H4+CH2CO 1.00E+12 0 0 REV/6.377E11 0.0E0 9.434E4/
799 C4H6+O=CH2O+C3H4-A 1.00E+12 0 0 REV/1.075E12 0.0E0 7.905E4/
800 C2H3+C2H4=C4H6+H 5.00E+11 0 7300 REV/1.0E13 0.0E0 4.7E3/
801 C4H8O1-2+OH=¿CH2O+C3H5-A+H2O 5.00E+12 0 0
802 C4H8O1-2+H=¿CH2O+C3H5-A+H2 5.00E+12 0 0
803 C4H8O1-2+O=¿CH2O+C3H5-A+OH 5.00E+12 0 0
804 C4H8O1-2+HO2=¿CH2O+C3H5-A+H2O2 1.00E+13 0 15000
805 C4H8O1-2+CH3O2=¿CH2O+C3H5-A+CH3O2H 1.00E+13 0 19000
806 C4H8O1-2+CH3=¿CH2O+C3H5-A+CH4 2.00E+11 0 10000
807 C4H8O1-3+OH=¿CH2O+C3H5-A+H2O 5.00E+12 0 0
808 C4H8O1-3+H=¿CH2O+C3H5-A+H2 5.00E+12 0 0
809 C4H8O1-3+O=¿CH2O+C3H5-A+OH 5.00E+12 0 0
810 C4H8O1-3+HO2=¿CH2O+C3H5-A+H2O2 1.00E+13 0 15000
811 C4H8O1-3+CH3O2=¿CH2O+C3H5-A+CH3O2H 1.00E+13 0 19000
812 C4H8O1-3+CH3=¿CH2O+C3H5-A+CH4 2.00E+11 0 10000
813 C4H8O1-4+OH=¿CH2O+C3H5-A+H2O 5.00E+12 0 0
814 C4H8O1-4+H=¿CH2O+C3H5-A+H2 5.00E+12 0 0
815 C4H8O1-4+O=¿CH2O+C3H5-A+OH 5.00E+12 0 0
816 C4H8O1-4+HO2=¿CH2O+C3H5-A+H2O2 1.00E+13 0 15000
817 C4H8O1-4+CH3O2=¿CH2O+C3H5-A+CH3O2H 1.00E+13 0 19000
818 C4H8O1-4+CH3=¿CH2O+C3H5-A+CH4 2.00E+11 0 10000
819 C4H8O2-3+OH=¿CH2O+C3H5-A+H2O 5.00E+12 0 0
820 C4H8O2-3+H=¿CH2O+C3H5-A+H2 5.00E+12 0 0
821 C4H8O2-3+O=¿CH2O+C3H5-A+OH 5.00E+12 0 0
822 C4H8O2-3+HO2=¿CH2O+C3H5-A+H2O2 1.00E+13 0 15000
823 C4H8O2-3+CH3O2=¿CH2O+C3H5-A+CH3O2H 1.00E+13 0 19000
824 C4H8O2-3+CH3=¿CH2O+C3H5-A+CH4 2.00E+11 0 10000
825 PC4H9O2=PC4H9+O2 2.85E+20 -1.642 35930 REV/4.52E12 0.0E0 0.0E0/
826 SC4H9O2=SC4H9+O2 4.33E+22 -2.216 38160 REV/7.54E12 0.0E0 0.0E0/
827 SC4H9O2+CH2O=SC4H9O2H+HCO 5.60E+12 0 13600 REV/8.0E11 0.0E0 1.0E4/
828 SC4H9O2+CH3CHO=SC4H9O2H+CH3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
829 SC4H9O2+HO2=SC4H9O2H+O2 1.75E+10 0 -3275 REV/3.861E13 -7.96E-1 3.362E4/
830 IC3H7O2+PC4H9=IC3H7O+PC4H9O 7.00E+12 0 -1000 REV/2.292E13 7.5E-2 2.909E4/
831 IC3H7O2+SC4H9=IC3H7O+SC4H9O 7.00E+12 0 -1000 REV/3.958E15 -4.18E-1 3.052E4/
269
Reaction A B E Auxiliary data

270
832 NC3H7O2+PC4H9=NC3H7O+PC4H9O 7.00E+12 0 -1000 REV/1.061E15 -6.11E-1 2.929E4/
833 NC3H7O2+SC4H9=NC3H7O+SC4H9O 7.00E+12 0 -1000 REV/1.832E17 -1.103E0 3.072E4/
834 2SC4H9O2=¿O2+2SC4H9O 1.40E+16 -1.61 1860
835 SC4H9O2+NC3H7O2=¿SC4H9O+NC3H7O+O2 1.40E+16 -1.61 1860
836 SC4H9O2+IC3H7O2=¿SC4H9O+IC3H7O+O2 1.40E+16 -1.61 1860
837 SC4H9O2+C2H5O2=¿SC4H9O+C2H5O+O2 1.40E+16 -1.61 1860
838 SC4H9O2+CH3O2=¿SC4H9O+CH3O+O2 1.40E+16 -1.61 1860
839 SC4H9O2+CH3CO3=¿SC4H9O+CH3CO2+O2 1.40E+16 -1.61 1860
840 PC4H9O2+HO2=¿PC4H9O+OH+O2 1.40E-14 -1.61 1860
841 SC4H9O2+HO2=¿SC4H9O+OH+O2 1.40E-14 -1.61 1860
842 H2+PC4H9O2=H+PC4H9O2H 3.01E+13 0 26030 REV/4.8E13 0.0E0 7.95E3/
843 H2+SC4H9O2=H+SC4H9O2H 3.01E+13 0 26030 REV/4.8E13 0.0E0 7.95E3/
844 C2H6+PC4H9O2=C2H5+PC4H9O2H 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
845 C2H6+SC4H9O2=C2H5+SC4H9O2H 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
846 PC4H9O2+C2H5CHO=PC4H9O2H+C2H5CO 2.00E+11 0 9500 REV/5.0E9 0.0E0 1.0E4/
847 SC4H9O2+C2H5CHO=SC4H9O2H+C2H5CO 2.00E+11 0 9500 REV/5.0E9 0.0E0 1.0E4/
848 SC4H9O2+CH3=SC4H9O+CH3O 7.00E+12 0 -1000 REV/1.181E13 6.7E-2 2.744E4/
849 SC4H9O2+C2H5=SC4H9O+C2H5O 7.00E+12 0 -1000 REV/1.679E15 -5.34E-1 2.985E4/
850 SC4H9O2+IC3H7=SC4H9O+IC3H7O 7.00E+12 0 -1000 REV/2.863E15 -3.68E-1 3.052E4/
851 SC4H9O2+NC3H7=SC4H9O+NC3H7O 7.00E+12 0 -1000 REV/1.695E15 -5.03E-1 2.852E4/
852 SC4H9O2+PC4H9=SC4H9O+PC4H9O 7.00E+12 0 -1000 REV/2.362E15 -5.54E-1 2.981E4/
853 SC4H9O2+SC4H9=2SC4H9O 7.00E+12 0 -1000 REV/4.08E17 -1.046E0 3.124E4/
854 SC4H9O2+C3H5-A=SC4H9O+C3H5O 7.00E+12 0 -1000 REV/4.373E11 2.49E-1 1.486E4/
855 PC4H9O2+CH2O=PC4H9O2H+HCO 5.60E+12 0 13600 REV/8.0E11 0.0E0 1.0E4/
856 PC4H9O2+CH3CHO=PC4H9O2H+CH3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
857 PC4H9O2+HO2=PC4H9O2H+O2 1.75E+10 0 -3275 REV/4.357E13 -8.13E-1 3.363E4/
858 C3H6+PC4H9O2=C3H5-A+PC4H9O2H 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
859 C3H6+SC4H9O2=C3H5-A+SC4H9O2H 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
860 C2H4+PC4H9O2=C2H3+PC4H9O2H 1.13E+13 0 30430 REV/3.0E12 0.0E0 1.15E4/
861 C2H4+SC4H9O2=C2H3+SC4H9O2H 1.13E+13 0 30430 REV/3.0E12 0.0E0 1.15E4/
862 CH3OH+PC4H9O2=CH2OH+PC4H9O2H 6.30E+12 0 19360 REV/1.0E9 0.0E0 1.0E4/
863 CH3OH+SC4H9O2=CH2OH+SC4H9O2H 6.30E+12 0 19360 REV/1.0E9 0.0E0 1.0E4/
864 C2H3CHO+PC4H9O2=C2H3CO+PC4H9O2H 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
865 C2H3CHO+SC4H9O2=C2H3CO+SC4H9O2H 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
866 CH4+PC4H9O2=CH3+PC4H9O2H 1.12E+13 0 24640 REV/7.43E11 0.0E0 5.5E3/
867 CH4+SC4H9O2=CH3+SC4H9O2H 1.12E+13 0 24640 REV/7.43E11 0.0E0 5.5E3/
868 C4H71-3+PC4H9O2=C4H7O+PC4H9O 7.00E+12 0 -1000 REV/1.192E14 -7.63E-1 1.771E4/
869 C4H71-3+SC4H9O2=C4H7O+SC4H9O 7.00E+12 0 -1000 REV/2.259E14 -6.81E-1 1.691E4/
870 H2O2+PC4H9O2=HO2+PC4H9O2H 2.40E+12 0 10000 REV/2.4E12 0.0E0 1.0E4/
871 H2O2+SC4H9O2=HO2+SC4H9O2H 2.40E+12 0 10000 REV/2.4E12 0.0E0 1.0E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
872 2PC4H9O2=¿O2+2PC4H9O 1.40E+16 -1.61 1860
873 PC4H9O2+SC4H9O2=¿PC4H9O+SC4H9O+O2 1.40E+16 -1.61 1860
874 PC4H9O2+NC3H7O2=¿PC4H9O+NC3H7O+O2 1.40E+16 -1.61 1860
875 PC4H9O2+IC3H7O2=¿PC4H9O+IC3H7O+O2 1.40E+16 -1.61 1860
876 PC4H9O2+C2H5O2=¿PC4H9O+C2H5O+O2 1.40E+16 -1.61 1860
877 PC4H9O2+CH3O2=¿PC4H9O+CH3O+O2 1.40E+16 -1.61 1860
878 PC4H9O2+CH3CO3=¿PC4H9O+CH3CO2+O2 1.40E+16 -1.61 1860
879 PC4H9O2+CH3=PC4H9O+CH3O 7.00E+12 0 -1000 REV/6.227E12 -1.5E-2 2.824E4/
880 PC4H9O2+C2H5=PC4H9O+C2H5O 7.00E+12 0 -1000 REV/8.854E14 -6.17E-1 3.065E4/
881 PC4H9O2+IC3H7=PC4H9O+IC3H7O 7.00E+12 0 -1000 REV/1.51E15 -4.5E-1 3.132E4/
882 PC4H9O2+NC3H7=PC4H9O+NC3H7O 7.00E+12 0 -1000 REV/8.939E14 -5.86E-1 2.932E4/
883 PC4H9O2+PC4H9=2PC4H9O 7.00E+12 0 -1000 REV/1.246E15 -6.36E-1 3.061E4/
884 PC4H9O2+SC4H9=PC4H9O+SC4H9O 7.00E+12 0 -1000 REV/2.152E17 -1.128E0 3.204E4/
885 PC4H9O2+C3H5-A=PC4H9O+C3H5O 7.00E+12 0 -1000 REV/2.307E11 1.66E-1 1.566E4/
886 PC4H9+HO2=PC4H9O+OH 7.00E+12 0 -1000 REV/8.669E15 -7.43E-1 2.661E4/
887 SC4H9+HO2=SC4H9O+OH 7.00E+12 0 -1000 REV/1.497E18 -1.235E0 2.804E4/
888 CH3O2+PC4H9=CH3O+PC4H9O 7.00E+12 0 -1000 REV/5.423E14 -4.44E-1 3.085E4/
889 CH3O2+SC4H9=CH3O+SC4H9O 7.00E+12 0 -1000 REV/9.365E16 -9.36E-1 3.228E4/
890 PC4H9O2H=PC4H9O+OH 1.50E+16 0 42500 REV/1.184E8 1.712E0 -2.732E3/
891 SC4H9O2H=SC4H9O+OH 9.45E+15 0 41600 REV/1.595E8 1.777E0 -4.422E3/
892 PC4H9O=NC3H7+CH2O 1.57E+21 -2.444 16840 REV/5.0E10 0.0E0 3.457E3/
893 SC4H9O=CH3+C2H5CHO 4.40E+16 -0.893 16530 REV/5.0E10 0.0E0 9.043E3/
894 SC4H9O=C2H5+CH3CHO 5.51E+22 -2.757 13980 REV/3.33E10 0.0E0 6.397E3/
895 PC4H9O2=C4H8OOH1-2 2.00E+11 0 26850 REV/5.597E8 3.39E-1 1.197E4/
896 PC4H9O2=C4H8OOH1-3 2.50E+10 0 20850 REV/3.231E9 -1.36E-1 7.871E3/
897 PC4H9O2=C4H8OOH1-4 4.69E+09 0 22350 REV/1.269E10 -5.23E-1 6.9E3/
898 SC4H9O2=C4H8OOH2-1 3.00E+11 0 29400 REV/7.442E11 -5.11E-1 1.394E4/
899 SC4H9O2=C4H8OOH2-3 2.00E+11 0 26850 REV/3.994E10 -1.96E-1 1.393E4/
900 SC4H9O2=C4H8OOH2-4 3.75E+10 0 24400 REV/9.302E10 -5.11E-1 8.944E3/
901 PC4H9O2=C4H8-1+HO2 5.04E+38 -8.11 40490 REV/1.599E30 -6.283E0 2.035E4/
902 SC4H9O2=C4H8-1+HO2 5.08E+42 -9.41 41490 REV/2.595E33 -7.347E0 1.661E4/
903 SC4H9O2=C4H8-2+HO2 5.04E+38 -8.11 40490 REV/1.618E29 -5.793E0 1.83E4/
904 C4H8OOH1-2=C4H8-1+HO2 8.83E+16 -1.488 16260 REV/1.0E11 0.0E0 1.1E4/
905 C4H8OOH2-1=C4H8-1+HO2 4.85E+20 -2.574 21180 REV/1.0E11 0.0E0 1.175E4/
906 C4H8OOH2-3=C4H8-2+HO2 6.22E+19 -2.513 21020 REV/1.0E11 0.0E0 1.175E4/
907 C4H8OOH1-2=¿C4H8O1-2+OH 6.00E+11 0 22000
908 C4H8OOH1-3=¿C4H8O1-3+OH 7.50E+10 0 15250
909 C4H8OOH1-4=¿C4H8O1-4+OH 9.38E+09 0 6000
910 C4H8OOH2-1=¿C4H8O1-2+OH 6.00E+11 0 22000
911 C4H8OOH2-3=¿C4H8O2-3+OH 6.00E+11 0 22000
271
Reaction A B E Auxiliary data

272
912 C4H8OOH2-4=¿C4H8O1-3+OH 7.50E+10 0 15250
913 C4H8OOH1-1=NC3H7CHO+OH 9.00E+14 0 1500 REV/1.733E8 1.89E0 3.347E4/
914 C4H8OOH2-2=C2H5COCH3+OH 9.00E+14 0 1500 REV/1.99E10 1.34E0 3.551E4/
915 C4H8OOH1-3=¿OH+CH2O+C3H6 6.64E+13 -0.16 29900
916 C4H8OOH2-4=¿OH+CH3CHO+C2H4 1.95E+18 -1.63 26790
917 C4H8OOH1-2O2=C4H8OOH1-2+O2 2.59E+24 -2.709 39860 REV/7.54E12 0.0E0 0.0E0/
918 C4H8OOH1-3O2=C4H8OOH1-3+O2 5.60E+22 -2.234 37960 REV/7.54E12 0.0E0 0.0E0/
919 C4H8OOH1-4O2=C4H8OOH1-4+O2 2.57E+20 -1.611 35680 REV/4.52E12 0.0E0 0.0E0/
920 C4H8OOH2-1O2=C4H8OOH2-1+O2 2.82E+20 -1.622 35700 REV/4.52E12 0.0E0 0.0E0/
921 C4H8OOH2-3O2=C4H8OOH2-3+O2 4.52E+22 -2.218 37880 REV/7.54E12 0.0E0 0.0E0/
922 C4H8OOH2-4O2=C4H8OOH2-4+O2 2.82E+20 -1.622 35700 REV/4.52E12 0.0E0 0.0E0/
923 C4H8OOH1-2O2=NC4KET12+OH 2.00E+11 0 26400 REV/3.199E4 1.323E0 4.893E4/
924 C4H8OOH1-3O2=NC4KET13+OH 2.50E+10 0 21400 REV/1.435E3 1.486E0 4.474E4/
925 C4H8OOH1-4O2=NC4KET14+OH 3.13E+09 0 19350 REV/1.726E2 1.494E0 4.269E4/
926 C4H8OOH2-1O2=NC4KET21+OH 1.00E+11 0 23850 REV/3.626E2 1.835E0 5.005E4/
927 C4H8OOH2-3O2=NC4KET23+OH 1.00E+11 0 23850 REV/1.739E3 1.731E0 4.913E4/
928 C4H8OOH2-4O2=NC4KET24+OH 1.25E+10 0 17850 REV/1.021E2 1.843E0 4.392E4/
929 NC4KET12=¿C2H5CHO+HCO+OH 1.05E+16 0 41600
930 NC4KET13=¿CH3CHO+CH2CHO+OH 1.05E+16 0 41600
931 NC4KET14=¿CH2CH2CHO+CH2O+OH 1.50E+16 0 42000
932 NC4KET21=¿CH2O+C2H5CO+OH 1.50E+16 0 42000
933 NC4KET23=¿CH3CHO+CH3CO+OH 1.05E+16 0 41600
934 NC4KET24=¿CH2O+CH3COCH2+OH 1.50E+16 0 42000
935 C2H5COCH3+OH=CH2CH2COCH3+H2O 7.55E+09 0.97 1586 REV/1.527E9 9.6E-1 2.102E4/
936 C2H5COCH3+OH=CH3CHCOCH3+H2O 8.45E+11 0 -228 REV/1.586E13 -2.3E-1 2.862E4/
937 C2H5COCH3+OH=C2H5COCH2+H2O 5.10E+11 0 1192 REV/6.628E13 -7.0E-1 2.767E4/
938 C2H5COCH3+HO2=CH2CH2COCH3+H2O2 2.38E+04 2.55 16490 REV/2.857E4 2.22E0 4.459E3/
939 C2H5COCH3+HO2=CH3CHCOCH3+H2O2 2.00E+11 0 8698 REV/2.229E13 -5.5E-1 6.08E3/
940 C2H5COCH3+HO2=C2H5COCH2+H2O2 2.38E+04 2.55 14690 REV/1.836E7 1.52E0 9.702E3/
941 C2H5COCH3+O=CH2CH2COCH3+OH 2.25E+13 0 7700 REV/4.616E11 -1.0E-2 9.882E3/
942 C2H5COCH3+O=CH3CHCOCH3+OH 3.07E+13 0 3400 REV/5.847E13 -2.3E-1 1.5E4/
943 C2H5COCH3+O=C2H5COCH2+OH 5.00E+12 0 5962 REV/6.592E13 -7.0E-1 1.519E4/
944 C2H5COCH3+H=CH2CH2COCH3+H2 9.16E+06 2 7700 REV/4.279E5 1.99E0 1.198E4/
945 C2H5COCH3+H=CH3CHCOCH3+H2 4.46E+06 2 3200 REV/1.934E7 1.77E0 1.689E4/
946 C2H5COCH3+H=C2H5COCH2+H2 9.30E+12 0 6357 REV/2.792E14 -7.0E-1 1.768E4/
947 C2H5COCH3+O2=CH2CH2COCH3+HO2 2.05E+13 0 51310 REV/1.74E10 3.2E-1 -1.895E3/
948 C2H5COCH3+O2=CH3CHCOCH3+HO2 1.55E+13 0 41970 REV/1.221E12 1.0E-1 -1.822E3/
949 C2H5COCH3+O2=C2H5COCH2+HO2 2.05E+13 0 49150 REV/1.118E13 -3.7E-1 2.988E3/
950 C2H5COCH3+CH3=CH2CH2COCH3+CH4 3.19E+01 3.17 7172 REV/3.893E1 3.16E0 1.193E4/
951 C2H5COCH3+CH3=CH3CHCOCH3+CH4 1.74E+00 3.46 3680 REV/1.971E2 3.23E0 1.785E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
952 C2H5COCH3+CH3=C2H5COCH2+CH4 1.62E+11 0 9630 REV/1.27E14 -7.0E-1 2.143E4/
953 C2H5COCH3+CH3O=CH2CH2COCH3+CH3OH 2.17E+11 0 6460 REV/4.887E9 2.0E-2 9.007E3/
954 C2H5COCH3+CH3O=CH3CHCOCH3+CH3OH 1.45E+11 0 2771 REV/3.031E11 -2.0E-1 1.473E4/
955 C2H5COCH3+CH3O=C2H5COCH2+CH3OH 2.17E+11 0 4660 REV/3.14E12 -6.8E-1 1.425E4/
956 C2H5COCH3+CH3O2=CH2CH2COCH3+CH3O2H 3.01E+12 0 19380 REV/7.525E12 -5.1E-1 4.329E3/
957 C2H5COCH3+CH3O2=CH3CHCOCH3+CH3O2H 2.00E+12 0 15250 REV/4.641E14 -7.3E-1 9.612E3/
958 C2H5COCH3+CH3O2=C2H5COCH2+CH3O2H 3.01E+12 0 17580 REV/4.836E15 -1.2E0 9.572E3/
959 C2H5COCH3+C2H3=CH2CH2COCH3+C2H4 5.00E+11 0 10400 REV/3.436E8 8.2E-1 1.403E4/
960 C2H5COCH3+C2H3=CH3CHCOCH3+C2H4 3.00E+11 0 3400 REV/1.914E10 6.0E-1 1.644E4/
961 C2H5COCH3+C2H3=C2H5COCH2+C2H4 6.15E+10 0 4278 REV/2.716E10 1.3E-1 1.495E4/
962 C2H5COCH3+C2H5=CH2CH2COCH3+C2H6 5.00E+10 0 13400 REV/9.549E12 -5.7E-1 1.399E4/
963 C2H5COCH3+C2H5=CH3CHCOCH3+C2H6 3.00E+10 0 8600 REV/5.319E14 -7.9E-1 1.86E4/
964 C2H5COCH3+C2H5=C2H5COCH2+C2H6 5.00E+10 0 11600 REV/6.136E15 -1.26E0 1.923E4/
965 CH3CHOOCOCH3=CH3CHCOCH3+O2 1.37E+17 -1.69 28460 REV/1.0E11 0.0E0 0.0E0/
966 CH3CHOOCOCH3=CH2CHOOHCOCH3 8.90E+12 0 29700 REV/4.703E13 -5.2E-1 1.696E4/
967 CH2CHOOHCOCH3=C2H3COCH3+HO2 2.03E+19 -2.35 14130 REV/7.0E10 0.0E0 7.8E3/
968 CH2CH2CHO=C2H4+HCO 3.13E+13 -0.52 24590 REV/1.5E11 0.0E0 8.3E3/
969 CH2CH2COCH3=C2H4+CH3CO 1.00E+14 0 18000 REV/1.0E11 0.0E0 0.0E0/
970 C2H5COCH2=CH2CO+C2H5 1.00E+14 0 35000 REV/1.0E11 0.0E0 0.0E0/
971 CH3CHCOCH3=C2H3COCH3+H 3.42E+16 -0.82 41770 REV/5.0E12 0.0E0 1.2E3/
972 CH3CHCOCH3=CH3CHCO+CH3 1.41E+15 -0.44 38340 REV/1.23E11 0.0E0 7.8E3/
973 NC3H7CHO+O2=NC3H7CO+HO2 1.20E+05 2.5 37560 REV/1.0E7 5.0E-1 4.0E3/
974 NC3H7CHO+OH=NC3H7CO+H2O 2.00E+06 1.8 -1300 REV/1.377E6 1.788E0 3.026E4/
975 NC3H7CHO+H=NC3H7CO+H2 4.14E+09 1.12 2320 REV/6.584E8 1.108E0 1.872E4/
976 NC3H7CHO+O=NC3H7CO+OH 5.94E+12 0 1868 REV/4.149E11 -1.2E-2 1.617E4/
977 NC3H7CHO+HO2=NC3H7CO+H2O2 4.09E+04 2.5 10200 REV/1.672E5 2.16E0 1.03E4/
978 NC3H7CHO+CH3=NC3H7CO+CH4 2.89E-03 4.62 3210 REV/1.93E-3 4.8E0 1.921E4/
979 NC3H7CHO+CH3O=NC3H7CO+CH3OH 1.00E+12 0 3300 REV/1.006E10 3.82E-1 1.796E4/
980 NC3H7CHO+CH3O2=NC3H7CO+CH3O2H 4.09E+04 2.5 10200 REV/3.371E5 1.986E0 7.067E3/
981 NC3H7CHO+OH=C3H6CHO-1+H2O 5.28E+09 0.97 1586 REV/1.091E9 9.62E-1 2.102E4/
982 NC3H7CHO+OH=C3H6CHO-2+H2O 4.68E+07 1.61 -35 REV/8.074E5 1.912E0 2.194E4/
983 NC3H7CHO+OH=C3H6CHO-3+H2O 5.52E+02 3.12 -1176 REV/1.266E4 2.868E0 2.771E4/
984 NC3H7CHO+HO2=C3H6CHO-1+H2O2 2.38E+04 2.55 16490 REV/2.919E4 2.214E0 4.467E3/
985 NC3H7CHO+HO2=C3H6CHO-2+H2O2 9.64E+03 2.6 13910 REV/9.872E2 2.574E0 4.424E3/
986 NC3H7CHO+HO2=C3H6CHO-3+H2O2 3.44E+12 0.05 17880 REV/4.685E14 -5.3E-1 1.53E4/
987 NC3H7CHO+CH3O2=C3H6CHO-1+CH3O2H 2.38E+04 2.55 16490 REV/5.885E4 2.04E0 1.239E3/
988 NC3H7CHO+CH3O2=C3H6CHO-2+CH3O2H 9.64E+03 2.6 13910 REV/1.99E3 2.4E0 1.196E3/
989 NC3H7CHO+CH3O2=C3H6CHO-3+CH3O2H 3.44E+12 0.05 17880 REV/9.445E14 -7.04E-1 1.207E4/
990 NC3H7CO=NC3H7+CO 1.00E+11 0 9600 REV/3.97E5 1.151E0 7.8E2/
991 C3H6CHO-1=C2H4+CH2CHO 7.40E+11 0 21970 REV/2.11E11 0.0E0 7.35E3/
273
Reaction A B E Auxiliary data

274
992 C3H6CHO-3=C2H5CHCO+H 8.43E+15 -0.6 40400 REV/5.0E12 0.0E0 1.2E3/
993 C3H6CHO-3=C2H3CHO+CH3 3.17E+14 -0.39 29900 REV/1.23E11 0.0E0 7.8E3/
994 C3H6CHO-2=SC3H5CHO+H 4.95E+12 -0.15 31300 REV/5.0E12 0.0E0 2.9E3/
995 C3H6CHO-2=C3H6+HCO 8.25E+12 -0.18 21900 REV/1.0E11 0.0E0 6.0E3/
996 C2H5CHCO+OH=¿NC3H7+CO2 3.73E+12 0 -1010
997 C2H5CHCO+H=¿NC3H7+CO 4.40E+12 0 1459
998 C2H5CHCO+O=¿C3H6+CO2 3.20E+12 0 -437
999 SC3H5CHO+OH=SC3H5CO+H2O 2.69E+10 0.76 -340 REV/4.831E10 7.7E-1 3.709E4/
1000 SC3H5CO=C3H5-S+CO 8.60E+15 0 23000 REV/1.0E11 0.0E0 6.0E3/
1001 SC3H5CHO+HO2=SC3H5CO+H2O2 1.00E+12 0 11920 REV/1.066E13 -3.2E-1 1.789E4/
1002 SC3H5CHO+CH3=SC3H5CO+CH4 3.98E+12 0 8700 REV/4.313E13 1.0E-2 3.146E4/
1003 SC3H5CHO+O=SC3H5CO+OH 7.18E+12 0 1389 REV/1.308E12 1.0E-2 2.157E4/
1004 SC3H5CHO+O2=SC3H5CO+HO2 4.00E+13 0 37600 REV/3.014E11 3.4E-1 2.394E3/
1005 SC3H5CHO+H=SC3H5CO+H2 2.60E+12 0 2600 REV/1.079E12 1.0E-2 2.488E4/
1006 C2H3COCH3+OH=¿CH3CHO+CH3CO 1.00E+11 0 0
1007 C2H3COCH3+OH=¿CH2CO+C2H3+H2O 5.10E+11 0 1192
1008 C2H3COCH3+HO2=¿CH2CHO+CH3CO+OH 6.03E+09 0 7949
1009 C2H3COCH3+HO2=¿CH2CO+C2H3+H2O2 8.50E+12 0 20460
1010 C2H3COCH3+CH3O2=¿CH2CHO+CH3CO+CH3O 3.97E+11 0 17050
1011 C2H3COCH3+CH3O2=¿CH2CO+C2H3+CH3O2H 3.01E+12 0 17580
1012 IC4H10(+M)=CH3+IC3H7(+M) 4.83E+16 0 79900 LOW/2.41E19 0.0E0 5.2576E4/ TROE/2.5E-1
7.5E2 1.0E-10 1.0E10/
1013 IC4H10=TC4H9+H 2.51E+98 -23.81 145300 REV/1.131E93 -2.2873E1 4.839E4/
1014 IC4H10=IC4H9+H 9.85E+95 -23.11 147600 REV/2.248E92 -2.2752E1 4.636E4/
1015 IC4H10+H=TC4H9+H2 1.81E+06 2.54 6756 REV/3.259E0 3.753E0 1.339E4/
1016 IC4H10+H=IC4H9+H2 6.02E+05 2.4 2583 REV/5.492E1 3.034E0 4.923E3/
1017 IC4H10+CH3=TC4H9+CH4 1.36E+00 3.65 7154 REV/2.234E-3 4.417E0 1.532E4/
1018 IC4H10+CH3=IC4H9+CH4 9.04E-01 3.46 4598 REV/7.524E-2 3.648E0 8.478E3/
1019 IC4H10+OH=TC4H9+H2O 2.93E+04 2.531 -1659 REV/5.584E-1 3.638E0 1.986E4/
1020 IC4H10+OH=IC4H9+H2O 6.65E+04 2.665 -168.9 REV/6.437E1 3.193E0 1.706E4/
1021 IC4H10+C2H5=IC4H9+C2H6 1.51E+12 0 10400 REV/3.2E11 0.0E0 1.23E4/
1022 IC4H10+C2H5=TC4H9+C2H6 1.00E+11 0 7900 REV/3.0E11 0.0E0 2.1E4/
1023 IC4H10+HO2=IC4H9+H2O2 1.22E+05 2.5 16690 REV/6.414E3 2.439E0 3.03E3/
1024 IC4H10+HO2=TC4H9+H2O2 1.50E+04 2.5 12260 REV/1.563E1 3.018E0 2.889E3/
1025 IC4H10+O=TC4H9+OH 1.97E+05 2.402 1150 REV/1.839E-1 3.596E0 6.368E3/
1026 IC4H10+O=IC4H9+OH 4.05E+07 2.034 5136 REV/1.916E3 2.649E0 6.064E3/
1027 IC4H10+CH3O=IC4H9+CH3OH 4.80E+11 0 7000 REV/9.402E8 4.21E-1 9.567E3/
1028 IC4H10+CH3O=TC4H9+CH3OH 1.90E+10 0 2800 REV/7.344E5 1.0E0 9.657E3/
1029 IC4H10+O2=IC4H9+HO2 1.80E+14 0 46000 REV/8.611E10 2.86E-1 -6.349E3/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
1030 IC4H10+O2=TC4H9+HO2 2.04E+13 0 41350 REV/1.926E8 8.65E-1 -6.709E3/
1031 IC4H10+CH3O2=IC4H9+CH3O2H 1.22E+05 2.5 16690 REV/1.248E5 1.994E0 1.435E3/
1032 IC4H10+C2H5O2=IC4H9+C2H5O2H 2.55E+13 0 20460 REV/2.623E13 -5.06E-1 5.208E3/
1033 IC4H10+CH3CO3=IC4H9+CH3CO3H 2.55E+13 0 20460 REV/2.147E12 -9.0E-3 9.57E3/
1034 IC4H10+NC3H7O2=IC4H9+NC3H7O2H 2.55E+13 0 20460 REV/2.678E13 -5.09E-1 5.0E3/
1035 IC4H10+IC3H7O2=IC4H9+IC3H7O2H 2.55E+13 0 20460 REV/2.688E13 -5.1E-1 5.0E3/
1036 IC4H10+IC4H9O2=IC4H9+IC4H9O2H 2.55E+13 0 20460 REV/2.679E13 -5.09E-1 5.01E3/
1037 IC4H10+TC4H9O2=IC4H9+TC4H9O2H 2.55E+13 0 20460 REV/2.684E13 -5.1E-1 5.01E3/
1038 IC4H10+O2CHO=IC4H9+HO2CHO 2.52E+13 0 20440 REV/1.006E0 3.06E0 2.375E3/
1039 IC4H10+O2CHO=TC4H9+HO2CHO 2.80E+12 0 16010 REV/1.893E1 2.72E0 4.678E3/
1040 IC4H10+SC4H9O2=IC4H9+SC4H9O2H 2.25E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
1041 IC4H10+SC4H9O2=TC4H9+SC4H9O2H 2.80E+12 0 16000 REV/5.0E11 0.0E0 6.5E3/
1042 IC4H10+PC4H9O2=IC4H9+PC4H9O2H 2.25E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
1043 IC4H10+PC4H9O2=TC4H9+PC4H9O2H 2.80E+12 0 16000 REV/5.0E11 0.0E0 6.5E3/
1044 IC4H10+CH3O2=TC4H9+CH3O2H 1.50E+04 2.5 12260 REV/3.04E2 2.573E0 1.294E3/
1045 IC4H10+C2H5O2=TC4H9+C2H5O2H 2.80E+12 0 16000 REV/1.0E11 0.0E0 1.0E4/
1046 IC4H10+CH3CO3=TC4H9+CH3CO3H 2.80E+12 0 16000 REV/1.0E11 0.0E0 1.0E4/
1047 IC4H10+NC3H7O2=TC4H9+NC3H7O2H 2.80E+12 0 16000 REV/1.0E11 0.0E0 1.0E4/
1048 IC4H10+IC3H7O2=TC4H9+IC3H7O2H 2.80E+12 0 16000 REV/1.0E11 0.0E0 1.0E4/
1049 IC4H10+IC4H9O2=TC4H9+IC4H9O2H 2.80E+12 0 16000 REV/1.0E11 0.0E0 1.0E4/
1050 IC4H10+TC4H9O2=TC4H9+TC4H9O2H 2.80E+12 0 16000 REV/1.0E11 0.0E0 1.0E4/
1051 IC4H10+IC4H9=TC4H9+IC4H10 2.50E+10 0 7900 REV/2.25E11 0.0E0 1.23E4/
1052 IC4H9+HO2=IC4H9O+OH 7.00E+12 0 -1000 REV/3.712E15 -6.58E-1 2.654E4/
1053 TC4H9+HO2=TC4H9O+OH 7.00E+12 0 -1000 REV/4.083E18 -1.329E0 2.865E4/
1054 CH3O2+IC4H9=CH3O+IC4H9O 7.00E+12 0 -1000 REV/2.322E14 -3.59E-1 3.078E4/
1055 CH3O2+TC4H9=CH3O+TC4H9O 7.00E+12 0 -1000 REV/2.554E17 -1.03E0 3.289E4/
1056 IC4H9=IC4H8+H 4.98E+32 -6.23 40070 REV/9.233E30 -5.844E0 9.025E3/
1057 IC4H9=C3H6+CH3 1.64E+37 -7.4 38670 REV/3.261E28 -5.503E0 1.483E4/
1058 TC4H9=H+IC4H8 4.65E+46 -9.83 55080 REV/4.368E46 -1.0023E1 1.974E4/
1059 TC4H9+O2=IC4H8+HO2 2.00E-18 0 5000 REV/2.0E-19 0.0E0 1.75E4/
1060 IC4H9+O2=IC4H8+HO2 2.00E-18 0 5000 REV/2.0E-19 0.0E0 1.75E4/
1061 NC3H7O2+IC4H9=NC3H7O+IC4H9O 7.00E+12 0 -1000 REV/4.543E14 -5.26E-1 2.922E4/
1062 NC3H7O2+TC4H9=NC3H7O+TC4H9O 7.00E+12 0 -1000 REV/4.997E17 -1.197E0 3.133E4/
1063 NC3H7O2+IC4H7=NC3H7O+IC4H7O 7.00E+12 0 -1000 REV/4.183E11 1.82E-1 1.35E4/
1064 SC4H9O2+IC4H9=SC4H9O+IC4H9O 7.00E+12 0 -1000 REV/1.011E15 -4.69E-1 2.974E4/
1065 SC4H9O2+TC4H9=SC4H9O+TC4H9O 7.00E+12 0 -1000 REV/1.112E18 -1.14E0 3.185E4/
1066 PC4H9O2+IC4H9=PC4H9O+IC4H9O 7.00E+12 0 -1000 REV/5.334E14 -5.51E-1 3.054E4/
1067 PC4H9O2+TC4H9=PC4H9O+TC4H9O 7.00E+12 0 -1000 REV/5.867E17 -1.222E0 3.265E4/
1068 PC4H9O2+IC4H7=PC4H9O+IC4H7O 7.00E+12 0 -1000 REV/4.912E11 1.57E-1 1.482E4/
1069 SC4H9O2+IC4H7=SC4H9O+IC4H7O 7.00E+12 0 -1000 REV/9.313E11 2.39E-1 1.402E4/
275
Reaction A B E Auxiliary data

276
1070 IC4H9O2=IC4H9+O2 6.64E+19 -1.575 36080 REV/2.26E12 0.0E0 0.0E0/
1071 TC4H9O2=TC4H9+O2 3.33E+24 -2.472 37870 REV/1.41E13 0.0E0 0.0E0/
1072 IC4H9O2+C4H10=IC4H9O2H+SC4H9 1.12E+13 0 17700 REV/5.0E11 0.0E0 6.5E3/
1073 TC4H9O2+C4H10=TC4H9O2H+SC4H9 1.12E+13 0 17700 REV/5.0E11 0.0E0 6.5E3/
1074 IC4H9O2+C4H10=IC4H9O2H+PC4H9 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
1075 TC4H9O2+C4H10=TC4H9O2H+PC4H9 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
1076 IC3H7O2+IC4H9=IC3H7O+IC4H9O 7.00E+12 0 -1000 REV/9.811E12 1.6E-1 2.902E4/
1077 IC3H7O2+TC4H9=IC3H7O+TC4H9O 7.00E+12 0 -1000 REV/1.079E16 -5.11E-1 3.113E4/
1078 IC3H7O2+IC4H7=IC3H7O+IC4H7O 7.00E+12 0 -1000 REV/9.035E9 8.68E-1 1.33E4/
1079 IC4H9O2+C3H6=IC4H9O2H+C3H5-A 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
1080 TC4H9O2+C3H6=TC4H9O2H+C3H5-A 3.24E+11 0 14900 REV/2.0E10 0.0E0 1.5E4/
1081 IC4H9O2+IC4H8=IC4H9O2H+IC4H7 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1082 TC4H9O2+IC4H8=TC4H9O2H+IC4H7 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1083 PC4H9O2+IC4H8=PC4H9O2H+IC4H7 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1084 SC4H9O2+IC4H8=SC4H9O2H+IC4H7 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1085 IC3H7O2+IC4H8=IC3H7O2H+IC4H7 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1086 NC3H7O2+IC4H8=NC3H7O2H+IC4H7 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1087 IC4H9O2+C4H8-1=IC4H9O2H+C4H71-3 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1088 TC4H9O2+C4H8-1=TC4H9O2H+C4H71-3 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1089 IC4H9O2+C4H8-2=IC4H9O2H+C4H71-3 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1090 TC4H9O2+C4H8-2=TC4H9O2H+C4H71-3 1.40E+12 0 14900 REV/3.16E11 0.0E0 1.3E4/
1091 CC4H8O+OH=¿CH2O+C3H5-A+H2O 5.00E+12 0 0
1092 CC4H8O+H=¿CH2O+C3H5-A+H2 3.51E+07 2 5000
1093 CC4H8O+O=¿CH2O+C3H5-A+OH 1.12E+14 0 5200
1094 CC4H8O+HO2=¿CH2O+C3H5-A+H2O2 1.00E+13 0 15000
1095 CC4H8O+CH3O2=¿CH2O+C3H5-A+CH3O2H 1.00E+13 0 19000
1096 CC4H8O+CH3=¿CH2O+C3H5-A+CH4 2.00E+11 0 10000
1097 C2H4+TC4H9O2=C2H3+TC4H9O2H 7.00E+11 0 17110 REV/1.0E11 0.0E0 1.0E4/
1098 TC4H9O2+CH4=TC4H9O2H+CH3 1.13E+13 0 20460 REV/7.5E8 0.0E0 1.28E3/
1099 H2+TC4H9O2=H+TC4H9O2H 3.01E+13 0 26030 REV/4.8E13 0.0E0 7.95E3/
1100 TC4H9O2+C2H6=TC4H9O2H+C2H5 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
1101 TC4H9O2+C3H8=TC4H9O2H+IC3H7 2.00E+12 0 17000 REV/5.0E11 0.0E0 6.5E3/
1102 TC4H9O2+C3H8=TC4H9O2H+NC3H7 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
1103 TC4H9O2+CH3OH=TC4H9O2H+CH2OH 6.30E+12 0 19360 REV/1.0E9 0.0E0 1.0E4/
1104 TC4H9O2+C2H5OH=TC4H9O2H+PC2H4OH 6.30E+12 0 19360 REV/3.061E12 0.0E0 2.21E4/
1105 TC4H9O2+C2H5OH=TC4H9O2H+SC2H4OH 4.20E+12 0 15000 REV/2.04E12 0.0E0 1.774E4/
1106 IC4H9O2+CH3CHO=IC4H9O2H+CH3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
1107 TC4H9O2+CH3CHO=TC4H9O2H+CH3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
1108 IC4H9O2+C2H3CHO=IC4H9O2H+C2H3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
1109 TC4H9O2+C2H3CHO=TC4H9O2H+C2H3CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
1110 IC4H9O2+C2H5CHO=IC4H9O2H+C2H5CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
1111 TC4H9O2+C2H5CHO=TC4H9O2H+C2H5CO 2.80E+12 0 13600 REV/1.0E12 0.0E0 1.0E4/
1112 IC4H9O2+HO2=IC4H9O2H+O2 1.75E+10 0 -3275 REV/3.843E13 -7.95E-1 3.362E4/
1113 TC4H9O2+HO2=TC4H9O2H+O2 1.75E+10 0 -3275 REV/3.85E13 -7.95E-1 3.362E4/
1114 IC4H9O2+H2O2=IC4H9O2H+HO2 2.40E+12 0 10000 REV/2.4E12 0.0E0 1.0E4/
1115 TC4H9O2+H2O2=TC4H9O2H+HO2 2.40E+12 0 10000 REV/2.4E12 0.0E0 1.0E4/
1116 IC4H9O2+CH2O=IC4H9O2H+HCO 1.30E+11 0 9000 REV/2.5E10 0.0E0 1.01E4/
1117 TC4H9O2+CH2O=TC4H9O2H+HCO 1.30E+11 0 9000 REV/2.5E10 0.0E0 1.01E4/
1118 IC4H9O2+CH3O2=¿IC4H9O+CH3O+O2 1.40E+16 -1.61 1860
1119 TC4H9O2+CH3O2=¿TC4H9O+CH3O+O2 1.40E+16 -1.61 1860
1120 IC4H9O2+C2H5O2=¿IC4H9O+C2H5O+O2 1.40E+16 -1.61 1860
1121 TC4H9O2+C2H5O2=¿TC4H9O+C2H5O+O2 1.40E+16 -1.61 1860
1122 IC4H9O2+CH3CO3=¿IC4H9O+CH3CO2+O2 1.40E+16 -1.61 1860
1123 TC4H9O2+CH3CO3=¿TC4H9O+CH3CO2+O2 1.40E+16 -1.61 1860
1124 2IC4H9O2=¿O2+2IC4H9O 1.40E+16 -1.61 1860
1125 IC4H9O2+TC4H9O2=¿IC4H9O+TC4H9O+O2 1.40E+16 -1.61 1860
1126 2TC4H9O2=¿O2+2TC4H9O 1.40E+16 -1.61 1860
1127 IC4H9O2+PC4H9O2=¿IC4H9O+PC4H9O+O2 1.40E+16 -1.61 1860
1128 TC4H9O2+PC4H9O2=¿TC4H9O+PC4H9O+O2 1.40E+16 -1.61 1860
1129 IC4H9O2+SC4H9O2=¿IC4H9O+SC4H9O+O2 1.40E+16 -1.61 1860
1130 TC4H9O2+SC4H9O2=¿TC4H9O+SC4H9O+O2 1.40E+16 -1.61 1860
1131 IC4H9O2+NC3H7O2=¿IC4H9O+NC3H7O+O2 1.40E+16 -1.61 1860
1132 TC4H9O2+NC3H7O2=¿TC4H9O+NC3H7O+O2 1.40E+16 -1.61 1860
1133 IC4H9O2+IC3H7O2=¿IC4H9O+IC3H7O+O2 1.40E+16 -1.61 1860
1134 TC4H9O2+IC3H7O2=¿TC4H9O+IC3H7O+O2 1.40E+16 -1.61 1860
1135 IC4H9O2+HO2=¿IC4H9O+OH+O2 1.40E+16 -1.61 1860
1136 TC4H9O2+HO2=¿TC4H9O+OH+O2 1.40E+16 -1.61 1860
1137 IC4H9O2+CH3=IC4H9O+CH3O 7.00E+12 0 -1000 REV/5.72E12 3.0E-3 2.802E4/
1138 IC4H9O2+C2H5=IC4H9O+C2H5O 7.00E+12 0 -1000 REV/8.133E14 -5.99E-1 3.043E4/
1139 IC4H9O2+IC3H7=IC4H9O+IC3H7O 7.00E+12 0 -1000 REV/1.387E15 -4.32E-1 3.11E4/
1140 IC4H9O2+NC3H7=IC4H9O+NC3H7O 7.00E+12 0 -1000 REV/8.212E14 -5.68E-1 2.91E4/
1141 IC4H9O2+PC4H9=IC4H9O+PC4H9O 7.00E+12 0 -1000 REV/1.145E15 -6.18E-1 3.039E4/
1142 IC4H9O2+SC4H9=IC4H9O+SC4H9O 7.00E+12 0 -1000 REV/1.977E17 -1.11E0 3.182E4/
1143 IC4H9O2+IC4H9=2IC4H9O 7.00E+12 0 -1000 REV/4.9E14 -5.33E-1 3.032E4/
1144 IC4H9O2+TC4H9=IC4H9O+TC4H9O 7.00E+12 0 -1000 REV/5.39E17 -1.204E0 3.243E4/
1145 IC4H9O2+C3H5-A=IC4H9O+C3H5O 7.00E+12 0 -1000 REV/2.119E11 1.84E-1 1.544E4/
1146 IC4H9O2+C4H71-3=IC4H9O+C4H7O 7.00E+12 0 -1000 REV/1.095E14 -7.45E-1 1.749E4/
1147 IC4H9O2+IC4H7=IC4H9O+IC4H7O 7.00E+12 0 -1000 REV/4.512E11 1.75E-1 1.46E4/
1148 TC4H9O2+CH3=TC4H9O+CH3O 7.00E+12 0 -1000 REV/7.824E11 2.29E-1 2.834E4/
1149 TC4H9O2+C2H5=TC4H9O+C2H5O 7.00E+12 0 -1000 REV/1.112E14 -3.72E-1 3.075E4/
277
Reaction A B E Auxiliary data

278
1150 TC4H9O2+IC3H7=TC4H9O+IC3H7O 7.00E+12 0 -1000 REV/1.898E14 -2.06E-1 3.142E4/
1151 TC4H9O2+NC3H7=TC4H9O+NC3H7O 7.00E+12 0 -1000 REV/1.123E14 -3.41E-1 2.942E4/
1152 TC4H9O2+PC4H9=TC4H9O+PC4H9O 7.00E+12 0 -1000 REV/1.566E14 -3.92E-1 3.071E4/
1153 TC4H9O2+SC4H9=TC4H9O+SC4H9O 7.00E+12 0 -1000 REV/2.704E16 -8.84E-1 3.214E4/
1154 TC4H9O2+IC4H9=TC4H9O+IC4H9O 7.00E+12 0 -1000 REV/6.703E13 -3.07E-1 3.064E4/
1155 TC4H9O2+TC4H9=2TC4H9O 7.00E+12 0 -1000 REV/7.373E16 -9.78E-1 3.275E4/
1156 TC4H9O2+C3H5-A=TC4H9O+C3H5O 7.00E+12 0 -1000 REV/2.898E10 4.11E-1 1.576E4/
1157 TC4H9O2+C4H71-3=TC4H9O+C4H7O 7.00E+12 0 -1000 REV/1.497E13 -5.19E-1 1.781E4/
1158 TC4H9O2+IC4H7=TC4H9O+IC4H7O 7.00E+12 0 -1000 REV/6.172E10 4.01E-1 1.492E4/
1159 IC4H9O2+C2H4=IC4H9O2H+C2H3 2.00E+11 0 6000 REV/2.0E10 0.0E0 8.0E3/
1160 IC4H9O2+CH4=IC4H9O2H+CH3 1.13E+13 0 20460 REV/7.5E8 0.0E0 1.28E3/
1161 H2+IC4H9O2=H+IC4H9O2H 3.01E+13 0 26030 REV/4.8E13 0.0E0 7.95E3/
1162 IC4H9O2+C2H6=IC4H9O2H+C2H5 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
1163 IC4H9O2+C3H8=IC4H9O2H+IC3H7 2.00E+12 0 17000 REV/5.0E11 0.0E0 6.5E3/
1164 IC4H9O2+C3H8=IC4H9O2H+NC3H7 1.70E+13 0 20460 REV/5.0E11 0.0E0 6.5E3/
1165 IC4H9O2+CH3OH=IC4H9O2H+CH2OH 6.30E+12 0 19360 REV/1.0E9 0.0E0 1.0E4/
1166 IC4H9O2+C2H5OH=IC4H9O2H+PC2H4OH 6.30E+12 0 19360 REV/3.061E12 0.0E0 2.21E4/
1167 IC4H9O2+C2H5OH=IC4H9O2H+SC2H4OH 4.20E+12 0 15000 REV/2.04E12 0.0E0 1.774E4/
1168 IC4H9O2H=IC4H9O+OH 1.50E+16 0 42500 REV/1.233E8 1.712E0 -2.942E3/
1169 TC4H9O2H=TC4H9O+OH 5.95E+15 0 42540 REV/6.677E6 1.939E0 -2.582E3/
1170 IC4H9O+HO2=IC3H7CHO+H2O2 1.00E+12 0 0 REV/1.38E13 -2.5E-1 7.213E4/
1171 IC4H9O+OH=IC3H7CHO+H2O 1.81E+13 0 0 REV/4.209E13 8.0E-2 1.036E5/
1172 IC4H9O+CH3=IC3H7CHO+CH4 2.40E+13 0 0 REV/3.368E14 8.0E-2 8.892E4/
1173 IC4H9O+O=IC3H7CHO+OH 6.00E+12 0 0 REV/1.415E12 8.0E-2 8.635E4/
1174 IC4H9O+H=IC3H7CHO+H2 1.99E+13 0 0 REV/1.069E13 8.0E-2 8.844E4/
1175 IC4H9O=IC3H7CHO+H 4.00E+14 0 21500 REV/1.139E10 1.08E0 2.5E3/
1176 IC4H9O=CH2O+IC3H7 2.00E+14 0 17500 REV/1.877E2 2.796E0 4.591E3/
1177 TC4H9O=CH3COCH3+CH3 9.56E+22 -2.548 18650 REV/1.5E11 0.0E0 1.19E4/
1178 IC4H9O+O2=IC3H7CHO+HO2 1.93E+11 0 1660 REV/7.026E8 5.64E-1 3.32E4/
1179 TC4H9O+O2=IC4H8O+HO2 8.10E+11 0 4700 REV/1.0E11 0.0E0 3.2E4/
1180 IC4H8O=IC3H7CHO 4.18E+13 0 52720 REV/1.392E10 5.5E-1 7.205E4/
1181 IC4H8O+OH=IC3H6CHO+H2O 1.25E+12 0 0 REV/9.609E5 1.069E0 3.656E4/
1182 IC4H8O+H=IC3H6CHO+H2 1.25E+12 0 0 REV/9.063E4 1.175E0 2.167E4/
1183 IC4H8O+HO2=IC3H6CHO+H2O2 2.50E+12 0 15000 REV/1.049E8 4.8E-1 2.066E4/
1184 IC4H8O+CH3O2=IC3H6CHO+CH3O2H 2.50E+12 0 19000 REV/2.04E9 3.6E-2 2.307E4/
1185 IC4H8O+CH3=IC3H6CHO+CH4 5.00E+10 0 10000 REV/3.307E6 7.29E-1 3.321E4/
1186 IC4H8O+O=IC3H6CHO+OH 1.25E+12 0 0 REV/4.704E4 1.156E0 2.025E4/
1187 IC3H7CHO=TC3H6CHO+H 2.30E+18 -0.91 92000 REV/2.0E14 0.0E0 0.0E0/
1188 IC3H7CHO=IC3H7+HCO 1.13E+17 -0.03 79760 REV/1.81E13 0.0E0 0.0E0/
1189 IC3H7CHO+HO2=IC3H7CO+H2O2 3.00E+12 0 11920 REV/7.987E11 -6.2E-2 1.036E4/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
1190 IC3H7CHO+HO2=TC3H6CHO+H2O2 8.00E+10 0 11920 REV/3.366E12 -4.2E-1 1.105E4/
1191 IC3H7CHO+CH3=IC3H7CO+CH4 3.98E+12 0 8700 REV/2.325E13 -6.0E-2 2.563E4/
1192 IC3H7CHO+O=IC3H7CO+OH 7.18E+12 0 1389 REV/7.052E11 -6.0E-2 1.574E4/
1193 IC3H7CHO+O2=IC3H7CO+HO2 4.00E+13 0 37600 REV/1.625E11 2.7E-1 -3.432E3/
1194 IC3H7CHO+OH=IC3H7CO+H2O 2.69E+10 0.76 -340 REV/1.164E10 7.5E-1 3.12E4/
1195 IC3H7CHO+OH=TC3H6CHO+H2O 1.68E+12 0 -781 REV/1.194E13 -9.0E-2 2.981E4/
1196 IC3H7CHO+H=IC3H7CO+H2 2.60E+12 0 2600 REV/1.196E9 6.33E-1 1.704E4/
1197 IC3H7CHO+OH=IC3H6CHO+H2O 3.12E+06 2 -298 REV/6.388E5 1.99E0 1.913E4/
1198 IC3H7CHO+HO2=IC3H6CHO+H2O2 2.74E+04 2.55 15500 REV/3.33E4 2.21E0 3.468E3/
1199 IC3H7CHO+CH3O2=IC3H6CHO+CH3O2H 4.76E+04 2.55 16490 REV/2.377E5 2.04E0 3.742E3/
1200 IC3H7CO=IC3H7+CO 2.87E+20 -2.194 14970 REV/1.5E11 0.0E0 4.81E3/
1201 IC3H6CHO=C3H6+HCO 1.03E+15 -0.62 23170 REV/1.0E11 0.0E0 7.8E3/
1202 IC3H6CHO=C2H3CHO+CH3 2.43E+13 -0.27 22470 REV/1.0E11 0.0E0 7.8E3/
1203 IC4H8OH=IC4H8+OH 9.23E+14 -0.562 28050 REV/9.93E11 0.0E0 -9.6E2/
1204 IO2C4H8OH=IC4H8OH+O2 1.92E+21 -2.347 35790 REV/1.2E11 0.0E0 -1.1E3/
1205 IO2C4H8OH=¿CH3COCH3+CH2O+OH 1.25E+10 0 18900
1206 IC4H9O2=IC4H8O2H-I 7.50E+10 0 24400 REV/1.815E11 -5.07E-1 8.946E3/
1207 TC4H9O2=TC4H8O2H-I 9.00E+11 0 29400 REV/2.027E9 1.23E-1 1.184E4/
1208 IC4H9O2=IC4H8O2H-T 1.00E+11 0 24100 REV/5.085E6 7.8E-1 1.078E4/
1209 IC4H9O2=IC4H8+HO2 4.53E+35 -7.22 39490 REV/5.991E26 -5.331E0 2.124E4/
1210 TC4H9O2=IC4H8+HO2 1.52E+43 -9.41 41490 REV/1.269E32 -7.203E0 1.716E4/
1211 IC4H8OOH-IO2=IC4H8O2H-I+O2 1.44E+20 -1.627 35690 REV/2.26E12 0.0E0 0.0E0/
1212 TC4H8OOH-IO2=TC4H8O2H-I+O2 5.17E+22 -2.257 37800 REV/2.26E12 0.0E0 0.0E0/
1213 IC4H8OOH-TO2=IC4H8O2H-T+O2 2.27E+27 -3.233 39640 REV/1.41E13 0.0E0 0.0E0/
1214 IC4H8OOH-IO2=IC4KETII+OH 2.50E+10 0 21400 REV/9.93E2 1.455E0 4.442E4/
1215 IC4H8OOH-TO2=IC4KETIT+OH 2.00E+11 0 26400 REV/9.551E4 1.24E0 4.873E4/
1216 IC4KETII=¿CH2O+C2H5CO+OH 1.50E+16 0 42000
1217 IC4KETIT=¿CH3COCH3+HCO+OH 9.50E+15 0 42540
1218 TC4H8O2H-I=IC4H8+HO2 1.07E+20 -2.085 19390 REV/3.97E11 0.0E0 1.262E4/
1219 IC4H8O2H-T=IC4H8+HO2 1.53E+16 -1.109 17560 REV/3.97E11 0.0E0 1.262E4/
1220 IC4H8O2H-I=¿CC4H8O+OH 7.50E+10 0 15250
1221 IC4H8O2H-T=¿IC4H8O+OH 6.00E+11 0 22000
1222 TC4H8O2H-I=¿IC4H8O+OH 6.00E+11 0 22000
1223 IC4H8O2H-I=¿OH+CH2O+C3H6 8.45E+15 -0.68 29170
1224 IC4H8=C3H5-T+CH3 1.92E+66 -14.22 128100 REV/1.561E56 -1.2293E1 2.61E4/
1225 IC4H8=IC4H7+H 3.07E+55 -11.49 114300 REV/1.428E55 -1.1738E1 2.64E4/
1226 IC4H8+H=C3H6+CH3 5.68E+33 -5.72 20000 REV/6.093E26 -4.209E0 2.72E4/
1227 IC4H8+H=IC4H7+H2 3.40E+05 2.5 2492 REV/6.32E4 2.528E0 1.816E4/
1228 IC4H8+O=¿CH2CO+2CH3 3.33E+07 1.76 76
1229 IC4H8+O=¿IC3H6CO+2H 1.66E+07 1.76 76
279
Reaction A B E Auxiliary data

280
1230 IC4H8+O=IC4H7+OH 1.21E+11 0.7 7633 REV/1.164E10 7.09E-1 2.189E4/
1231 IC4H8+CH3=IC4H7+CH4 4.42E+00 3.5 5675 REV/7.495E2 3.082E0 2.289E4/
1232 IC4H8+HO2=IC4H7+H2O2 1.93E+04 2.6 13910 REV/2.073E6 1.933E0 1.358E4/
1233 IC4H8+O2CHO=IC4H7+HO2CHO 1.93E+04 2.6 13910 REV/6.514E-7 4.9E0 -3.468E3/
1234 IC4H8+O2=IC4H7+HO2 6.00E+12 0 39900 REV/5.848E12 -3.2E-1 8.83E2/
1235 IC4H8+C3H5-A=IC4H7+C3H6 7.94E+11 0 20500 REV/4.4E20 -1.33E0 6.061E4/
1236 IC4H8+C3H5-S=IC4H7+C3H6 7.94E+11 0 20500 REV/5.592E20 -1.27E0 8.217E4/
1237 IC4H8+C3H5-T=IC4H7+C3H6 7.94E+11 0 20500 REV/5.592E20 -1.27E0 8.017E4/
1238 IC4H8+OH=IC4H7+H2O 5.20E+06 2 -298 REV/1.025E7 1.922E0 3.027E4/
1239 IC4H8+O=IC3H7+HCO 1.58E+07 1.76 -1216 REV/4.538E0 3.06E0 2.169E4/
1240 IC4H8+CH3O2=IC4H7+CH3O2H 1.93E+04 2.6 13910 REV/4.034E7 1.488E0 1.199E4/
1241 IC4H8+HO2=IC4H8O+OH 1.29E+12 0 13340 REV/1.0E12 0.0E0 7.5E3/
1242 IC4H7+O2=IC3H5CHO+OH 2.47E+13 -0.45 23020 REV/3.372E13 -5.77E-1 7.301E4/
1243 IC4H7+O2=CH3COCH2+CH2O 7.14E+15 -1.21 21050 REV/1.7E12 -4.07E-1 8.825E4/
1244 IC4H7+O2=¿C3H4-A+CH2O+OH 7.29E+29 -5.71 21450
1245 IC4H7+O=IC3H5CHO+H 6.03E+13 0 0 REV/2.844E16 -5.19E-1 6.673E4/
1246 IC4H7=C3H4-A+CH3 1.23E+47 -9.74 74260 REV/1.649E38 -7.768E0 2.254E4/
1247 CH3O2+IC4H7=CH3O+IC4H7O 7.00E+12 0 -1000 REV/2.138E11 3.49E-1 1.506E4/
1248 IC4H7+HO2=IC4H7O+OH 7.00E+12 0 -1000 REV/3.418E12 5.0E-2 1.082E4/
1249 IC4H7O=C3H5-T+CH2O 2.93E+21 -2.391 35590 REV/1.0E11 0.0E0 1.26E4/
1250 IC4H7O=IC4H6OH 1.39E+11 0 15600 REV/4.233E11 -1.64E-1 3.167E4/
1251 IC4H7O=IC3H5CHO+H 5.00E+13 0 29100 REV/6.67E13 -1.05E-1 1.841E4/
1252 IC4H6OH+H2=IC4H7OH+H 2.16E+04 2.38 18990 REV/5.614E2 2.98E0 1.399E3/
1253 IC4H6OH+HO2=IC4H7OH+O2 5.57E+13 -0.315 862 REV/6.0E13 0.0E0 3.99E4/
1254 IC4H6OH+CH2O=IC4H7OH+HCO 6.30E+08 1.9 18190 REV/2.101E7 2.153E0 1.773E4/
1255 IC4H6OH+IC4H8=IC4H7OH+IC4H7 4.70E+02 3.3 19840 REV/2.814E-1 3.9E0 6.521E3/
1256 IC4H7OH=IC4H6OH+H 4.90E+16 -0.4 89850 REV/1.0E14 0.0E0 0.0E0/
1257 IC4H7OH+HO2=IC4H6OH+H2O2 7.64E+03 2.712 13930 REV/7.83E5 2.05E0 1.358E4/
1258 IC4H6OH=C3H4-A+CH2OH 7.24E+19 -1.859 57050 REV/1.0E11 0.0E0 9.2E3/
1259 IC4H7O+O2=IC3H5CHO+HO2 3.00E+10 0 1649 REV/6.312E10 -1.4E-1 3.898E4/
1260 IC4H7O+HO2=IC3H5CHO+H2O2 3.00E+11 0 0 REV/8.93E14 -8.0E-1 7.85E4/
1261 IC4H7O+CH3=IC3H5CHO+CH4 2.40E+13 0 0 REV/7.261E16 -4.7E-1 9.529E4/
1262 IC4H7O+O=IC3H5CHO+OH 6.00E+12 0 0 REV/3.052E14 -4.7E-1 9.272E4/
1263 IC4H7O+OH=IC3H5CHO+H2O 1.81E+13 0 0 REV/9.076E15 -4.7E-1 1.1E5/
1264 IC4H7O+H=IC3H5CHO+H2 1.99E+13 0 0 REV/2.305E15 -4.7E-1 9.481E4/
1265 IC3H5CHO+OH=IC3H5CO+H2O 2.69E+10 0.76 -340 REV/4.4E10 7.8E-1 3.608E4/
1266 IC3H5CHO+HO2=IC3H5CO+H2O2 1.00E+12 0 11920 REV/9.709E12 -3.1E-1 1.688E4/
1267 IC3H5CHO+CH3=IC3H5CO+CH4 3.98E+12 0 8700 REV/3.928E13 2.0E-2 3.045E4/
1268 IC3H5CHO+O=IC3H5CO+OH 7.18E+12 0 1389 REV/1.191E12 2.0E-2 2.056E4/
1269 IC3H5CHO+O2=IC3H5CO+HO2 2.00E+13 0 40700 REV/1.824E11 3.11E-1 5.337E3/
Chemical kinetic mechanism
Reaction A B E Auxiliary data
1270 IC3H5CHO+H=IC3H5CO+H2 2.60E+12 0 2600 REV/9.822E11 2.0E-2 2.387E4/
1271 IC3H5CO=C3H5-T+CO 1.28E+20 -1.89 34460 REV/1.51E11 0.0E0 4.809E3/
1272 TC3H6CHO+HO2=TC3H6OCHO+OH 9.64E+12 0 0 REV/2.018E17 -1.2E0 2.101E4/
1273 TC3H6OCHO=CH3COCH3+HCO 3.98E+13 0 9700 REV/2.173E8 8.0E-1 1.424E4/
1274 TC3H6CHO=IC3H5CHO+H 1.33E+14 0.01 39340 REV/1.3E13 0.0E0 1.2E3/
1275 TC3H6CHO=IC3H6CO+H 4.09E+14 -0.072 42410 REV/1.3E13 0.0E0 4.8E3/
1276 TC3H6CHO+H2=IC3H7CHO+H 2.16E+05 2.38 18990 REV/1.319E5 2.47E0 3.55E3/
1277 IC4H7OOH=IC4H7O+OH 6.40E+15 0 45550 REV/1.0E11 0.0E0 0.0E0/
1278 IC4H7OH=IC4H7O+H 5.97E+16 -0.56 105900 REV/4.0E13 0.0E0 0.0E0/
1279 IC4H8OH=IC4H7OH+H 1.71E+12 0.277 38850 REV/1.0E13 0.0E0 1.2E3/
1280 IC4H7O+H2=IC4H7OH+H 9.05E+06 2 17830 REV/7.16E5 2.44E0 1.631E4/
1281 IC4H7OH=IC4H7+OH 7.31E+16 -0.41 79700 REV/3.0E13 0.0E0 0.0E0/
1282 IC4H7O+CH2O=IC4H7OH+HCO 1.15E+11 0 1280 REV/3.02E11 0.0E0 1.816E4/
1283 TC3H6CHO+CH2O=IC3H7CHO+HCO 2.52E+08 1.9 18190 REV/1.229E7 1.99E0 1.742E4/
1284 TC3H6CHO+IC4H8=IC3H7CHO+IC4H7 4.70E+02 3.3 19840 REV/6.613E0 3.39E0 8.672E3/
1285 IC3H6CO+OH=IC3H7+CO2 1.73E+12 0 -1010 REV/2.577E14 -4.3E-1 5.548E4/
1286 TC3H6OHCHO=TC3H6CHO+OH 9.99E+20 -1.46 87480 REV/5.0E13 0.0E0 0.0E0/
1287 TC3H6OHCHO=TC3H6OH+HCO 5.16E+23 -1.9 76850 REV/1.81E13 0.0E0 0.0E0/
1288 TC3H6OH=CH3COCH3+H 5.00E+13 0 21860 REV/1.0E12 0.0E0 0.0E0/
1289 TC3H6OH=IC3H5OH+H 6.20E+15 -0.66 40340 REV/1.3E13 0.0E0 1.56E3/
1290 IC3H5OH=C3H5-T+OH 7.37E+19 -0.94 109100 REV/5.0E13 0.0E0 0.0E0/
1291 TC3H6O2CHO=TC3H6CHO+O2 2.46E+25 -4.065 27080 REV/1.99E17 -2.1E0 0.0E0/
1292 TC3H6O2CHO=IC3H5O2HCHO 6.00E+11 0 29880 REV/3.014E12 -5.16E-1 1.711E4/
1293 TC3H6O2CHO=TC3H6O2HCO 1.00E+11 0 25750 REV/1.628E12 -5.16E-1 2.51E4/
1294 IC3H5O2HCHO=IC3H5CHO+HO2 8.94E+20 -2.44 15030 REV/2.23E11 0.0E0 1.06E4/
1295 TC3H6O2HCO=¿CH3COCH3+CO+OH 4.24E+18 -1.43 4800
1296 TC3H6OH+O2=CH3COCH3+HO2 2.23E+13 0 0 REV/6.529E17 -1.19E0 2.561E4/
1297 IC3H6CO+OH=TC3H6OH+CO 2.00E+12 0 -1010 REV/1.009E9 9.0E-1 3.154E4/
1298 TC3H6CHO+O2=IC3H5CHO+HO2 2.73E-19 0 7240 REV/1.39E11 -2.0E-1 1.731E4/
1299 TC3H6CHO+O2=¿CH3COCH3+CO+OH 3.62E-20 0 0
1300 TC3H6CHO+HO2=IC3H7CHO+O2 3.68E+12 0 1310 REV/1.236E14 -2.4E-1 4.335E4/
1301 TC3H6CHO+CH3=IC3H5CHO+CH4 3.01E+12 -0.32 -131 REV/2.207E15 -8.5E-1 6.79E4/
1302 TC4H8CHO=IC3H5CHO+CH3 1.00E+13 0 26290 REV/2.23E11 0.0E0 1.06E4/
1303 TC4H8CHO=IC4H8+HCO 8.52E+12 0 20090 REV/1.0E11 0.0E0 6.0E3/
1304 O2C4H8CHO=TC4H8CHO+O2 1.52E+19 -1.44 34510 REV/2.0E12 0.0E0 0.0E0/
1305 O2C4H8CHO=O2HC4H8CO 2.16E+11 0 15360 REV/1.173E13 -6.8E-1 1.488E4/
1306 O2HC4H8CO=IC4H8O2H-T+CO 3.30E+22 -2.72 11760 REV/1.5E11 0.0E0 4.809E3/
1307 IC4H7O+IC4H8=IC4H7OH+IC4H7 2.70E+11 0 4000 REV/1.0E10 0.0E0 9.0E3/
1308 IC4H6OH+HO2=¿CH2CCH2OH+CH2O+OH 1.45E+13 0 0
1309 IC4H8+CH2CCH2OH=IC4H7+C3H5OH 7.94E+11 0 20500 REV/2.75E11 -5.0E-2 2.847E4/
281
Reaction A B E Auxiliary data

282
1310 C3H5OH+HO2=CH2CCH2OH+H2O2 1.76E+09 0.28 22590 REV/3.01E9 0.0E0 2.583E3/
1311 C3H5OH+OH=CH2CCH2OH+H2O 5.06E+12 0 5960 REV/1.457E12 5.0E-2 1.741E4/
1312 C3H5OH+H=CH2CCH2OH+H2 3.90E+05 2.5 5821 REV/2.594E4 2.55E0 2.121E3/
1313 C3H5OH+O2=CH2CCH2OH+HO2 4.00E+13 0 60690 REV/4.833E10 3.8E-1 -4.92E2/
1314 C3H5OH+CH3=CH2CCH2OH+CH4 2.40E+11 0 8030 REV/4.17E11 5.0E-2 4.81E3/
1315 IC4H7OH=CH2CCH2OH+CH3 1.25E+20 -0.98 98570 REV/3.0E13 0.0E0 0.0E0/
1316 C3H5OH=CH2CCH2OH+H 2.84E+19 -1.05 111100 REV/1.0E14 0.0E0 0.0E0/
1317 CH2CCH2OH+O2=¿CH2OH+CO+CH2O 4.34E+12 0 0
1318 CH2CCH2OH=C2H2+CH2OH 2.16E+40 -8.31 45110 REV/1.61E40 -8.58E0 2.033E4/
1319 CH2CCH2OH=C3H4-A+OH 6.70E+16 -1.11 42580 REV/8.5E12 0.0E0 2.0E3/
1320 N2+M=2N+M 1.89E+18 -0.85 224950 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1321 N+O+M=NO+M 7.60E+14 -0.1 -1770 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1322 N2+O=N+NO 1.00E+14 0 75490
1323 N+NO2=N2O+O 1.80E+12 0 0
1324 N+O2=NO+O 6.40E+09 1 6280
1325 N+OH=NO+H 3.80E+13 0 0
1326 NH+O2=HNO+O 4.60E+05 2 6500
1327 NH+O2=NO+OH 1.30E+06 1.5 100
1328 NH+OH=HNO+H 2.00E+13 0 0
1329 NH+OH=N+H2O 5.00E+11 0.5 2000
1330 NH+OH=NO+H2 2.00E+13 0 0
1331 NH+H=N+H2 3.00E+13 0 0
1332 NH+O=NO+H 9.20E+13 0 0
1333 NH+N=N2+H 3.00E+13 0 0
1334 2NH=N2+2H 2.54E+13 0 0
1335 NH+NO=N2O+H 4.30E+14 -0.5 0
1336 NH+NO=N2+OH 2.20E+13 -0.23 0
1337 NH+NO2=N2O+OH 1.00E+13 0 0
1338 NH2+O2=HNO+OH 4.50E+12 0 25000
1339 NH2+O=NH+OH 6.80E+12 0 0
1340 NH2+O=H2+NO 1.30E+08 1.025 -627
1341 NH2+O=HNO+H 6.60E+14 -0.5 0
1342 NH2+OH=NH+H2O 4.00E+06 2 1000
Chemical kinetic mechanism
Reaction A B E Auxiliary data
1343 NH2+H=NH+H2 4.00E+13 0 3650
1344 NH2+NO=NNH+OH 2.80E+13 -0.55 0
1345 NH2+NO=N2+H2O 5.00E+12 0 0
1346 NH2+NO=N2O+H2 3.90E+13 0 20160
1347 NH2+N=N2+2H 7.20E+13 0 0
1348 NH2+NH=N2H2+H 5.00E+13 0 0
1349 2NH2=N2H2+H2 8.50E+11 0 0
1350 2NH2=NH3+NH 4.00E+13 0 10000
1351 NH2+HO2=H2NO+OH 5.00E+13 0 0
1352 NH2+HO2=NH3+O2 1.00E+13 0 0
1353 NH2+NO2=N2O+H2O 3.20E+18 -2.2 0
1354 NH3+M=NH+H2+M 6.31E+14 0 93400 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1355 NH3+M=NH2+H+M 2.51E+16 0 93800 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1356 NH3+OH=NH2+H2O 2.04E+06 2.04 566
1357 NH3+H=NH2+H2 6.36E+05 2.39 10171
1358 NH3+O=NH2+OH 9.40E+06 1.94 6460
1359 NH3+HO2=NH2+H2O2 3.00E+11 0 22000
1360 NNH+M=N2+H+M 2.00E+14 0 20000 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1361 NNH+NO=N2+HNO 5.00E+13 0 0
1362 NNH+H=N2+H2 1.00E+14 0 0
1363 NNH+OH=N2+H2O 5.00E+13 0 0
1364 NNH+NH2=N2+NH3 5.00E+13 0 0
1365 NNH+NH=N2+NH2 5.00E+13 0 0
1366 NNH+O=N2O+H 1.00E+14 0 0
1367 NNH+O=N2+OH 8.00E+13 0 0
1368 NNH+O=NH+NO 5.00E+13 0 0
1369 NNH+O2=N2+HO2 2.00E+14 0 0
1370 NNH+O2=N2+O2+H 5.00E+13 0 0
1371 HNO+O=NO+OH 1.00E+13 0 0
1372 HNO+OH=NO+H2O 3.60E+13 0 0
283
Reaction A B E Auxiliary data

284
1373 HNO+H=NO+H2 4.40E+11 0.72 650
1374 HNO+NH2=NO+NH3 2.00E+13 0 1000
1375 HNO+N=NO+NH 1.00E+13 0 1990
1376 HNO+NO2=NO+HONO 6.00E+11 0 2000
1377 2HNO=N2O+H2O 3.95E+12 0 5000
1378 HNO+NO=N2O+OH 2.00E+12 0 26000
1379 N2H2+M=NNH+H+M 5.00E+16 0 50000 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1380 N2H2+H=NNH+H2 5.00E+13 0 1000
1381 N2H2+O=NH2+NO 1.00E+13 0 1000
1382 N2H2+O=NNH+OH 2.00E+13 0 1000
1383 N2H2+OH=NNH+H2O 1.00E+13 0 1000
1384 N2H2+NO=N2O+NH2 3.00E+12 0 0
1385 N2H2+NH=NNH+NH2 1.00E+13 0 1000
1386 N2H2+NH2=NH3+NNH 1.00E+13 0 1000
1387 NO+HO2=NO2+OH 2.10E+12 0 -480
1388 NO+OH+M=HONO+M 5.10E+23 -2.51 -68 Third body: H2O /5.0/
1389 NO+H+M=HNO+M 4.00E+20 -1.75 0 Third body: H2O /4.1/ Third body: H2 /1.25/
Third body: N2 /1.0/
1390 NO+HCO=HNO+CO 7.20E+12 0 0
1391 NO2+H2=HONO+H 2.40E+13 0 29000
1392 NO2+H=NO+OH 1.00E+14 0 362
1393 NO2+O=NO+O2 1.00E+13 0 600
1394 NO2+M=NO+O+M 1.10E+16 0 66000 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1395 2NO2=2NO+O2 2.00E+12 0 26825
1396 NO2+HO2=HONO+O2 6.30E+08 1.25 5000
1397 NO2+HCO=CO+NO+OH 9.45E+12 0 -430
1398 NO2+HCO=H+CO2+NO 5.55E+12 0 -430
1399 NO2+CO=NO+CO2 2.19E+13 0 29200
1400 NO2+NO=N2O+O2 1.00E+12 0 60000
1401 NO2+CH2=NO+CH2O 5.90E+13 0 0
1402 NO2+CH=NO+HCO 1.00E+14 0 0
1403 N2O+OH=N2+HO2 2.00E+12 0 40000
1404 N2O+H=N2+OH 2.08E-06 5.557 1820
Chemical kinetic mechanism
Reaction A B E Auxiliary data
1405 N2O+M=N2+O+M 3.00E+14 0 55500 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1406 N2O+O=N2+O2 1.40E+12 0 10800
1407 N2O+O=2NO 2.90E+13 0 23150
1408 N2O+N=N2+NO 1.00E+13 0 19870
1409 N2O+CO=N2+CO2 2.70E+11 0 20237
1410 N2O+NO=NO2+N2 5.30E+05 2.23 46281
1411 N2O+CH=HCN+NO 1.90E+13 0 -511
1412 CH+NO=HCN+O 4.80E+13 0 0
1413 CH+NO=HCO+N 3.40E+13 0 0
1414 CH+NO=NCO+H 1.90E+13 0 0
1415 CH2+NO=HCNO+H 1.30E+12 0 -378
1416 CH2+NO=HCN+OH 2.20E+12 0 -378
1417 CH3+NO=HCN+H2O 1.50E-01 3.523 3950
1418 CH3+NO=H2CN+OH 1.50E-01 3.523 3950
1419 C2H3+NO=C2H2+HNO 1.00E+12 0 1000
1420 C2H+NO=CN+HCO 2.10E+13 0 0
1421 HCCO+NO=HCNO+CO 4.89E+12 0.059 -584
1422 HCCO+NO=HCN+CO2 1.64E+13 -0.267 -707
1423 HCNO+H=HCN+OH 5.70E+12 0 0
1424 HCNO+O=NO+HCO 7.00E+13 0 0
1425 HCNO+OH=NO+CH2O 2.00E+13 0 0
1426 HOCN+OH=NCO+H2O 6.40E+05 2 2560
1427 HOCN+O=NCO+OH 1.50E+04 2.64 4000
1428 HOCN+H=HNCO+H 2.00E+07 2 2000
1429 HOCN+H=NCO+H2 2.00E+07 2 2000
1430 HNCO+M=NH+CO+M 1.10E+16 0 86000
1431 HNCO+HO2=NCO+H2O2 3.00E+11 0 29000
1432 HNCO+O2=HNO+CO2 1.00E+12 0 35000
1433 HNCO+NH2=NH3+NCO 5.00E+12 0 6200
1434 HNCO+NH=NH2+NCO 3.00E+13 0 23700
1435 HNCO+H=NH2+CO 2.20E+07 1.7 3800
1436 HNCO+CN=NCO+HCN 2.50E+12 0 0
1437 HNCO+O=NCO+OH 2.20E+06 2.11 11430
1438 HNCO+O=CO2+NH 9.65E+07 1.41 8520
1439 HNCO+O=HNO+CO 1.50E+08 1.57 44012
1440 HNCO+OH=NCO+H2O 4.20E+05 2 2560
285

1441 HCN+OH=CN+H2O 3.20E+06 1.83 10300


Reaction A B E Auxiliary data

286
1442 HCN+OH=HOCN+H 5.85E+04 2.4 12500
1443 HCN+OH=HNCO+H 4.00E-03 4 1000
1444 HCN+OH=NH2+CO 7.83E-04 4 4000
1445 HCN+O=NCO+H 1.38E+04 2.64 4980
1446 HCN+O=NH+CO 3.45E+03 2.64 4980
1447 HCN+O=CN+OH 2.70E+09 1.58 29200
1448 CN+OH=NH+CO 6.00E+12 0 0
1449 CN+OH=HNCO 6.00E+12 0 0
1450 CN+OH=NCO+H 6.00E+13 0 0
1451 CN+NO=NCO+N 9.64E+13 0 42100
1452 CN+NO=CO+N2 9.64E+13 0 42100
1453 CN+HNO=HCN+NO 1.80E+13 0 0
1454 CN+HONO=HCN+NO2 1.21E+13 0 0
1455 CN+H2=HCN+H 3.00E+05 2.45 2237
1456 CN+O=CO+N 7.70E+13 0 0
1457 CN+O2=NCO+O 7.50E+12 0 -389
1458 CN+HCN=C2N2+H 1.50E+07 1.71 1530
1459 CN+NO2=NCO+NO 5.30E+15 -0.75 344
1460 CN+NO2=CO+N2O 4.90E+14 -0.75 344
1461 CN+NO2=N2+CO2 3.70E+14 -0.75 344
1462 CN+N2O=NCO+N2 1.00E+13 0 0
1463 CN+N2O=NCN+NO 3.80E+03 2.6 3700
1464 CN+CO2=NCO+CO 3.70E+06 2.16 26900
1465 CN+CH2O=HCN+HCO 4.22E+13 0 0
1466 C2N2+O=NCO+CN 4.57E+12 0 8880
1467 C2N2+OH=HOCN+CN 1.86E+11 0 2900
1468 NCO+H=NH+CO 5.00E+13 0 0
1469 NCO+CH2O=HNCO+HCO 6.02E+12 0 0
1470 NCO+HCO=HNCO+CO 3.62E+13 0 0
1471 NCO+NO=N2O+CO 5.23E+17 -1.73 763
1472 NCO+NO=CO2+N2 4.11E+17 -1.73 763
1473 NCO+NO2=CO+2NO 1.39E+13 0 0
1474 NCO+NO2=CO2+N2O 5.40E+12 0 0
1475 NCO+HNO=HNCO+NO 1.80E+13 0 0
1476 NCO+HONO=HNCO+NO2 3.60E+12 0 0
1477 NCO+N2O=N2+NO+CO 9.00E+13 0 27800
1478 2NCO=N2+2CO 1.80E+13 0 0
1479 NCO+O=NO+CO 2.00E+13 0 0
1480 NCO+N=N2+CO 2.00E+13 0 0
1481 NCO+OH=HCO+NO 5.00E+12 0 15000
Chemical kinetic mechanism
Reaction A B E Auxiliary data
1482 NCO+M=N+CO+M 3.10E+16 -0.5 48000 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1483 NCO+H2=HNCO+H 7.60E+02 3 4000
1484 NCO+O2=NO+CO2 2.00E+12 0 20000
1485 NCO+CN=NCN+CO 1.80E+13 0 0
1486 NCN+H=HCN+N 1.00E+14 0 0
1487 NCN+O=CN+NO 1.00E+14 0 0
1488 NCN+OH=HCN+NO 5.00E+13 0 0
1489 NCN+O2=NO+NCO 1.00E+13 0 0
1490 H2NO+O=NH2+O2 2.00E+14 0 0
1491 H2NO+M=H2+NO+M 7.83E+27 -4.29 60300 Third body: H2O /10.0/
1492 H2NO+M=HNO+H+M 2.80E+24 -2.83 64915 Third body: H2O /10.0/
1493 H2NO+HO2=HNO+H2O2 2.90E+04 2.69 -1600
1494 H2NO+O2=HNO+HO2 3.00E+12 0 25000
1495 H2NO+H=HNO+H2 3.00E+07 2 2000
1496 H2NO+H=NH2+OH 5.00E+13 0 0
1497 H2NO+O=HNO+OH 3.00E+07 2 2000
1498 H2NO+OH=HNO+H2O 2.00E+07 2 1000
1499 H2NO+NO=2HNO 2.00E+04 2 13000
1500 H2NO+NH2=HNO+NH3 3.00E+12 0 1000
1501 HONO+O=NO2+OH 1.20E+13 0 6000
1502 HONO+OH=NO2+H2O 1.30E+10 1 135
1503 CH2+N2=HCN+NH 1.00E+13 0 74000
1504 CH+N2=HCN+N 3.00E+11 0 13600
1505 CH2+N=HCN+H 5.00E+13 0 0
1506 CH+N=CN+H 1.30E+13 0 0
1507 CO2+N=CO+NO 1.90E+11 0 3400
1508 CH3+N=H2CN+H 3.00E+13 0 0
1509 C2H3+N=HCN+CH2 2.00E+13 0 0
1510 C3H3+N=HCN+C2H2 1.00E+13 0 0
1511 H2CN+N=N2+CH2 2.00E+13 0 0
1512 H2CN+M=HCN+H+M 3.00E+14 0 22000 Third body: H2O /16.25/ Third body: CO /1.875/
Third body: CO2 /3.75/ Third body: CH4 /16.25/
Third body: C2H6 /16.25/
1513 CH2O+NO2=HCO+HONO 8.00E+02 2.77 13730
1514 HCO+HNO=CH2O+NO 6.00E+11 0 2000
287
Reaction A B E Auxiliary data

288
1515 2HONO=H2O+NO+NO2 3.49E-01 3.64 12140
1516 NCO+CH4=HNCO+CH3 1.00E+13 0 8126
1517 NCO+C2H6=HNCO+C2H5 1.45E-09 6.89 2916
1518 CH3O+HNO=CH3OH+NO 3.16E+13 0 0
1519 CH3+NO(+M)=CH3NO(+M) 2.20E+11 0.6 0 LOW/2.06E27 -3.5E0 0.0E0/
1520 CH3NO2(+M)=CH3+NO2(+M) 1.80E+16 0 58500 LOW/1.3E17 0.0E0 4.2E4/ TROE/1.83E-1 1.0E-30
1.0E30/
1521 CH3+NO2=CH3O+NO 1.51E+13 0 0
1522 CH3O+NO2=CH2O+HONO 6.02E+12 0 2285
1523 CH3O+NO2(+M)=CH3ONO2(+M) 1.20E+13 0 0 LOW/1.4E30 -4.5E0 0.0E0/
1524 CH3O+NO=CH2O+HNO 1.30E+14 -0.7 0
1525 CH3O2+NO=CH3O+NO2 5.50E+11 0 -1192
1526 CH3NO2+O=H2CNO2+OH 1.51E+13 0 5350
1527 CH3NO2+OH=H2CNO2+H2O 1.49E+13 0 5740
1528 CH3NO2+OH=CH3OH+NO2 2.00E+10 0 -1000
1529 CH3NO2+O2=H2CNO2+HO2 2.00E+13 0 57000
1530 CH3NO2+CH2=CH3+H2CNO2 6.50E+12 0 7900
1531 CH3NO2+H=H2CNO2+H2 7.50E+12 0 10000
1532 CH3NO2+H=HONO+CH3 3.27E+12 0 3730
1533 CH3NO2+H=CH3NO+OH 1.40E+12 0 3730
1534 CH3NO2+CH3=H2CNO2+CH4 7.08E+11 0 11140
1535 CH3NO2+CH3O=H2CNO2+CH3OH 3.00E+11 0 7000
1536 CH3NO2+C2H5=H2CNO2+C2H6 3.00E+11 0 11700
1537 CH3NO2+HO2=H2CNO2+H2O2 1.50E+11 0 15000
1538 H2CNO2=CH2O+NO 1.00E+13 0 36000
1539 H2CNO2+HONO=CH3NO2+NO2 1.00E+12 0 0
1540 H2CNO2+H=CH3+NO2 5.00E+13 0 0
1541 H2CNO2+O=CH2O+NO2 5.00E+13 0 0
1542 H2CNO2+OH=CH2OH+NO2 1.00E+13 0 0
1543 H2CNO2+OH=CH2O+HONO 1.00E+13 0 0
1544 CH3O+NO(+M)=CH3ONO(+M) 1.21E+13 0 -332 LOW/2.7E27 -3.5E0 0.0E0/
1545 CH3NO2=CH3ONO 2.90E+14 0 67000
1546 C2H5+NO2=C2H5O+NO 1.00E+13 0 0
1547 C2H5O2+NO=C2H5O+NO2 3.00E+12 0 -358
1548 HO2+NO+M=HONO2+M 2.23E-12 -3.5 2200
1549 NO2+OH(+M)=HONO2(+M) 3.61E+13 0 0 LOW/1.44E25 -2.9E0 0.0E0/
1550 HONO2+OH=H2O+NO3 1.03E+10 0 -1240
1551 NO3+OH=HO2+NO2 1.38E+13 0 0
1552 NO3+O=O2+NO2 1.02E+13 0 0
Chemical kinetic mechanism
Reaction A B E Auxiliary data
1553 NO3+H=NO2+OH 6.00E+13 0 0
1554 NO3+HO2=O2+HONO2 5.60E+11 0 0
1555 NO3+HO2=O2+OH+NO2 2.00E+12 0 0
1556 2NO3=O2+2NO2 5.12E+11 0 4840
1557 NO3+M=O2+NO+M 2.05E+08 1 12122
1558 NO3+NO2=NO2+NO+O2 2.35E+10 0 2960
1559 NO3+NO=2NO2 1.08E+13 0 -219
1560 NO2+O(+M)=NO3(+M) 1.20E+13 0 0 LOW/2.94E21 -2.0E0 0.0E0/
1561 NO2+O3=NO3+O2 7.23E+10 0 4870
1562 O+O2+M=O3+M 1.88E+21 -2.8 0
1563 O+O3=2O2 4.80E+12 0 4090
1564 H+O3=OH+O2 8.43E+13 0 950
1565 OH+O3=HO2+O2 1.14E+12 0 2000
1566 NO+O3=NO2+O2 1.08E+12 0 2720
1567 HO2+O3=OH+2O2 8.43E+09 0 1200
1568 CH3+O3=CH3O+O2 1.57E+12 0 0
1569 CH3+HONO=CH4+NO2 8.10E+05 1.87 5504
1570 C2H5+HONO=C2H6+NO2 8.10E+05 1.87 5504
1571 C2H3+HONO=C2H4+NO2 8.10E+05 1.87 5504
1572 CH3O+HONO=CH3OH+NO2 8.10E+05 1.87 5504
1573 CH3O2H+H=CH3O+H2O 8.20E+10 0 1860
1574 CH3O2H+O=CH3O2+OH 1.00E+12 0 3000
1575 CH3O2H+OH=CH3O2+H2O 1.80E+12 0 -378
289
Minerva Access is the Institutional Repository of The University of Melbourne

Author/s:
Morganti, Kai J.

Title:
A study of the knock limits of liquefied petroleum gas (LPG) in spark-ignition engines

Date:
2013

Citation:
Morganti, K. J. (2013). A study of the knock limits of liquefied petroleum gas (LPG) in spark-
ignition engines. PhD thesis, Department of Mechanical Engineering, The University of
Melbourne.

Persistent Link:
http://hdl.handle.net/11343/38535

File Description:
A study of the knock limits of liquefied petroleum gas (LPG) in spark-ignition engines

Terms and Conditions:


Terms and Conditions: Copyright in works deposited in Minerva Access is retained by the
copyright owner. The work may not be altered without permission from the copyright owner.
Readers may only download, print and save electronic copies of whole works for their own
personal non-commercial use. Any use that exceeds these limits requires permission from
the copyright owner. Attribution is essential when quoting or paraphrasing from these works.

Potrebbero piacerti anche