Vous êtes sur la page 1sur 268

Chemical Kinetic Modeling of Biofuel Combustion

by
Subram Maniam Sarathy
A thesis submitted in conformity with the requirements
for the degree of Doctor of Philosophy
Graduate Department of Chemical Engineering and Applied Chemistry
University of Toronto
Copyright c 2010 by Subram Maniam Sarathy
Abstract
Chemical Kinetic Modeling of Biofuel Combustion
Subram Maniam Sarathy
Doctor of Philosophy
Graduate Department of Chemical Engineering and Applied Chemistry
University of Toronto
2010
Bioalcohols, such as bioethanol and biobutanol, are suitable replacements for gaso-
line, while biodiesel can replace petroleum diesel. Improving biofuel engine performance
requires understanding its fundamental combustion properties and the pathways of com-
bustion. This studys contribution is experimentally validated chemical kinetic com-
bustion mechanisms for biobutanol and biodiesel. Fundamental combustion data and
chemical kinetic mechanisms are presented and discussed to improve our understanding
of biofuel combustion.
The net environmental impact of biobutanol (i.e., n-butanol) has not been studied
extensively, so this study rst assesses the sustainability of n-butanol derived from corn.
The results indicate that technical advances in fuel production are required before com-
mercializing biobutanol. The primary contribution of this research is new experimental
data and a novel chemical kinetic mechanism for n-butanol combustion. The results
indicate that under the given experimental conditions, n-butanol is consumed primarily
via abstraction of hydrogen atoms to produce fuel radical molecules, which subsequently
decompose to smaller hydrocarbon and oxygenated species. The hydroxyl moiety in
n-butanol results in the direct production of the oxygenated species such as butanal,
acetaldehyde, and formaldehyde. The formation of these compounds sequesters carbon
from forming soot precursors, but they may introduce other adverse environmental and
health eects.
Biodiesel is a mixture of long chain fatty acid methyl esters derived from fats and
ii
oils. This research study presents high quality experimental data for one large fatty acid
methyl ester, methyl decanoate, and models its combustion using an improved skeletal
mechanism. The results indicate that methyl decanoate is consumed via abstraction
of hydrogen atoms to produce fuel radicals, which ultimately lead to the production of
alkenes. The ester moiety in methyl decanoate leads to the formation of low molecular
weight oxygenated compounds such as carbon monoxide, formaldehyde, and ketene.
The study concludes that the oxygenated molecules in biofuels follow similar combus-
tion pathways to the hydrocarbons in petroleum fuels. The oxygenated moietys ability to
sequester carbon from forming soot precursors is highlighted. However, the direct forma-
tion of oxygenated hydrocarbons warrants further investigation into the environmental
and health impacts of practical biofuel combustion systems.
iii
Dedication
Sabbo pajjalito loko, sabbo loko pakampito
The entire universe is nothing but combustion and vibration
To my forefathers and their wives
Sivarajan the Medical Doctor
Viswanathan the Industrialist
Subramaniam the Entrepreneur
Radhakrishnan the Doctor of Engineering
Gurumoorthy the Renunciate
To all beings
May you be happy and peaceful.
May you enjoy good health and harmony.
iv
Acknowledgements
I oer greatest thanks to my wife, Nimisha Rajawat. I would have never been able to
complete this work without her contributions of patience, compassion, encouragement,
love, and delicious food. Thanks for being their and bringing Panya into my life.
Endless gratitude to Professor Murray Thomson for his guidance and support in my
academic and professional endeavours. Thanks for being a wise guide, a passionate
leader, and a good friend.
Much appreciation to Professor Kirk and Professor Wallace for being members on my
PhD supervisory committee. I also thank Professor Mims and Professor Tran for serving
on my examination committee. It was an honour to have Professor Seshadri from UC
San Diego serving as my external examiner. His insights into my work were invaluable.
Thanks to Professor Phillipe Dagaut, Professor Heather MacLean, Dr. Yimin Zhang,
Dr. William Pitz, Professor Tianfeng Lu, and my other coauthors for their eorts.
Gratefulness to all my colleagues in the Combustion Research Group, specically Dr.
Qingan Zhang, Dr. Seth Dworkin, Dr. Salvador Rego, Professor Zhenyu Wen, Dr.
Jerome Thiebaud, Richard Mills, Phil Geddis, Parham Zabeti, Meghdad Saaripour,
Tim Chan, and Coleman Yeung. My contemporary, Tom Tzanetakis, deserves a special
recognition for sharing his knowledge of both practical engineering systems and funda-
mental scientic theory. He is the smartest and most humble engineer that I know.
Acknowledgement to NSERC and AUTO21 for funding my research and studies.
My father, Roger Sarathy, deserves special a acknowledgement for proofreading this
dissertation. Also, thanks for always encouraging me to ask questions and to search for
answers.
Besides my wife, two other women deserve special acknowledegment: my mother, Saraswathi
Sarathy, and her sister (i.e., my aunt), Durga Krishnan. Their regular phone calls and
barrage of internet messages always kept me on my toes. Thanks for being there and
v
making sure I never slacked o.
Much respect to Dr. Siva Sarathy, Professor Brinda Sarathy, Brihas Sarathy, Maggie
Decarie, Professor Karthick Ramakrishan, Chef Siddarth Deepak, Dr. David Miller, Wei
Lun Huang, and all my family and friends. Words cannot express my gratitude towards
you, so here is a loving smile :)
Thanks to my Dhamma family at the Ontario Vipassana Centre for supporting my
journey towards fullling the ten paramis (i.e., perfect qualities): generosity, moral-
ity, renunciation, wisdom, energy, patience, truthfulness, determination, loving kindness,
and equanimity. The development of these paramis has provided immeasurable benets
in my research.
vi
Contents
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii
Dedication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii
List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvi
Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii
Statement of Co-Authorship and Copyright . . . . . . . . . . . . . . . . . . . xx
I Background and Methods 1
1 Introduction 2
1.1 Research Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Dissertation Objectives and Layout . . . . . . . . . . . . . . . . . . . . . 5
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2 Background 8
2.1 Reciprocating ICEs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.1 SI Engines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.2 CI Engines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Fuel Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Combustion Emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3 Modeling Combustion Chemistry 15
3.1 Chemical Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2 Computer Simulations for Mechanism Validation . . . . . . . . . . . . . . 16
vii
3.2.1 Governing Equations for Chemically Reacting Flows . . . . . . . . 17
Conservation of Mass . . . . . . . . . . . . . . . . . . . . . . . . . 18
Conservation of Momentum . . . . . . . . . . . . . . . . . . . . . 18
Conservation of Species . . . . . . . . . . . . . . . . . . . . . . . . 19
Conservation of Energy . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3 Solving the Governing Equations . . . . . . . . . . . . . . . . . . . . . . 21
3.3.1 Chemical Kinetic Database . . . . . . . . . . . . . . . . . . . . . 22
3.3.2 Thermochemical Database . . . . . . . . . . . . . . . . . . . . . . 24
3.3.3 Transport Database . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.4 Developing Chemical Kinetic Mechanisms . . . . . . . . . . . . . . . . . 28
3.4.1 Mechanisms for Hydrocarbon Fuels . . . . . . . . . . . . . . . . . 30
Combustion of Alkanes . . . . . . . . . . . . . . . . . . . . . . . . 30
Combustion of Alkenes . . . . . . . . . . . . . . . . . . . . . . . . 33
Mechanism of Soot Formation . . . . . . . . . . . . . . . . . . . . 34
3.4.2 Determining Rate Coecients . . . . . . . . . . . . . . . . . . . . 36
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4 Experimental Apparatus and Analytical Methodology 43
4.1 Opposed-ow Diusion Burner Setup . . . . . . . . . . . . . . . . . . . . 44
4.2 Fuel Preparation and Vaporization . . . . . . . . . . . . . . . . . . . . . 45
4.3 Supply of Fuel and Oxidizer Streams . . . . . . . . . . . . . . . . . . . . 50
4.4 Reynolds Number and Strain Rate Calculations . . . . . . . . . . . . . . 52
4.5 Gas Sampling System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.5.1 Sampling Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.5.2 Sampling Procedure . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.6 Analytical Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.6.1 Non-Dispersive Infrared Analysis . . . . . . . . . . . . . . . . . . 58
CO and CO
2
Measurements . . . . . . . . . . . . . . . . . . . . . 58
4.6.2 Gas Chromatography . . . . . . . . . . . . . . . . . . . . . . . . . 59
GC Carrier Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Injection System . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
GC Measurement Procedures . . . . . . . . . . . . . . . . . . . . 61
GC Calibration Procedure . . . . . . . . . . . . . . . . . . . . . . 63
4.6.3 Temperature Measurement . . . . . . . . . . . . . . . . . . . . . . 66
viii
Correction for Radiation Losses . . . . . . . . . . . . . . . . . . . 68
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
II Biobutanol 71
5 Background 72
5.1 Biobutanol History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.2 Biobutanol Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.3 Biobutanol Fuel Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6 LCA of Biobutanol for use in Transportation 80
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
6.1.1 Bioethanol LCA . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
Land Use Change and Food Security Issues . . . . . . . . . . . . . 82
6.1.2 Biobutanol LCA . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.2 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.2.1 Data sources and uncertainties . . . . . . . . . . . . . . . . . . . . 84
6.2.2 Corn ethanol and butanol production . . . . . . . . . . . . . . . . 85
6.2.3 Post Production Life Cycle Activities . . . . . . . . . . . . . . . . 87
6.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.3.1 WTP fossil energy use . . . . . . . . . . . . . . . . . . . . . . . . 89
6.3.2 WTP petroleum use . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.3.3 WTP GHG emissions . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.3.4 WTW Fossil Energy Use and GHG Emissions . . . . . . . . . . . 91
6.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.4.1 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
7 Chemical Kinetic Modeling of Butanol Combustion 100
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
7.1.1 Engine Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
7.1.2 Combustion Chemistry Studies . . . . . . . . . . . . . . . . . . . 101
7.2 Experimental Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
ix
7.2.1 Opposed-ow Diusion Flame . . . . . . . . . . . . . . . . . . . . 102
7.2.2 Jet Stirred Reactor . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.2.3 Laminar Flame Speed Setup . . . . . . . . . . . . . . . . . . . . . 104
7.3 Computational Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
7.3.1 Chemical Kinetic Mechanism . . . . . . . . . . . . . . . . . . . . 106
7.3.2 Thermochemical Data . . . . . . . . . . . . . . . . . . . . . . . . 109
7.3.3 Transport Properties . . . . . . . . . . . . . . . . . . . . . . . . . 109
7.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.4.1 Jet Stirred Reactor . . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.4.2 Opposed-ow Diusion Flame . . . . . . . . . . . . . . . . . . . . 119
7.4.3 Laminar Flame Speed . . . . . . . . . . . . . . . . . . . . . . . . 125
7.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
7.6 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Supplemental Material . . . . . . . . . . . . . . . . . . . . . . . . 130
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
III Biodiesel 137
8 Background 138
Biodiesel Sustainability . . . . . . . . . . . . . . . . . . . . . . . . 139
8.1 Biodiesel Fuel Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
8.1.1 Biodiesel Production . . . . . . . . . . . . . . . . . . . . . . . . . 141
8.2 Biodiesel Fuel Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
8.3 Biodiesel Exhaust Emissions . . . . . . . . . . . . . . . . . . . . . . . . . 144
8.3.1 CO, THC, and Oxygenate Emissions . . . . . . . . . . . . . . . . 145
8.3.2 PM and NO
a
emissions . . . . . . . . . . . . . . . . . . . . . . . . 146
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
9 Chemical Kinetic Modeling of Biodiesel Combustion 153
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
9.2 Mechanisms for Short Chain Methyl Esters . . . . . . . . . . . . . . . . . 154
9.3 Mechanisms for Long Chain Methyl Esters . . . . . . . . . . . . . . . . . 157
9.3.1 Mechanisms for Methyl Decanoate . . . . . . . . . . . . . . . . . 159
9.4 Background Summary and Research Motivation . . . . . . . . . . . . . . 160
x
9.5 Experimental Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
9.6 Computational Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
9.6.1 Chemical Kinetic Mechanism . . . . . . . . . . . . . . . . . . . . 162
Modied Detailed Chemical Kinetic Mechanism . . . . . . . . . . 163
Skeletal Chemical Kinetic Mechanism . . . . . . . . . . . . . . . . 168
9.6.2 Thermochemical Data . . . . . . . . . . . . . . . . . . . . . . . . 172
9.6.3 Transport Properties . . . . . . . . . . . . . . . . . . . . . . . . . 172
9.7 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
9.7.1 Opposed-ow Diusion Flame . . . . . . . . . . . . . . . . . . . . 174
9.7.2 Temperature, Fuel, and Hydrocarbon Species . . . . . . . . . . . 175
9.7.3 Oxygenated Species . . . . . . . . . . . . . . . . . . . . . . . . . . 179
The Fate of the Ester Moiety . . . . . . . . . . . . . . . . . . . . 183
9.7.4 Jet Stirred Reactor . . . . . . . . . . . . . . . . . . . . . . . . . . 185
9.8 Conclusions and Recommendations . . . . . . . . . . . . . . . . . . . . . 186
Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
IV Closing 196
10 Scientic Contribution 197
Appendix A 199
Appendix B 200
Appendix C 240
xi
List of Tables
2.1 Selected characteristics of SI and CI engines . . . . . . . . . . . . . . . . 9
3.1 Relative magnitudes of rate constants for H abstraction from dierent CH
bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.1 Physical Chemical Properties of the Fuels Used . . . . . . . . . . . . . . 46
4.2 Experimental conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.3 Dual Column GC Method Parameters . . . . . . . . . . . . . . . . . . . 63
4.4 GC Method Parameters for Fatty Acid Methyl Esters . . . . . . . . . . . 64
4.5 Measured FID relative molar response factors for organic molecules . . . 65
5.1 Selected fuel properties of butanol, ethanol, and gasoline . . . . . . . . . 77
6.1 Data for estimating energy use and GHG emissions for corn butanol pro-
duction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
7.1 Chemical structures of species during the oxidation of n-butanol . . . . . 107
7.2 Comparison of maximum measured and predicted product concentrations
for n-butanol in the JSR . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
7.3 Comparison of maximum measured species in n-butanol (C
4
H
9
OH) and
n-butane (C
4
H
10
) opposed-ow diusion ames. . . . . . . . . . . . . . . 123
8.1 Typical fatty acid composition (wt%) of various biodiesel feedstock [17] . 142
8.2 Properties of common biodiesel fuels and pure FAME . . . . . . . . . . . 145
8.3 Percent of publications that report changes in emissions for biodiesel [23]. 146
9.1 Experimentally and Empirically Determined Polarizabilities (

A
3
) for FAME174
9.2 maximum Measured and Predicted Concentration of Oxygenated Species 181
xii
List of Figures
3.1 A typical phase diagram showing critical point . . . . . . . . . . . . . . . 26
3.2 Flowchart for developing and validating chemical kinetic mechanisms . . 29
3.3 Decay of isopropyl radical . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.4 Addition of O radical to ethene . . . . . . . . . . . . . . . . . . . . . . . 34
3.5 Chemical structures and bond dissociation energies for alcohols and alka-
nes [24] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.6 General mechanism for soot formation from Glassman [27] . . . . . . . . 38
4.1 Diagram of burner port . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2 Photograph of burner setup . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.3 Schematic of the mixing chamber . . . . . . . . . . . . . . . . . . . . . . 49
4.4 Schematic of the machined ange plate. . . . . . . . . . . . . . . . . . . . 49
4.5 Vapor pressure curves for fatty acid methyl esters . . . . . . . . . . . . . 51
4.6 Schematic of microprobe . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.7 Schematic of microprobe and burner setup . . . . . . . . . . . . . . . . . 57
4.8 Schematic of dual column GC Setup . . . . . . . . . . . . . . . . . . . . 62
4.9 Schematic of the permeation tube setup . . . . . . . . . . . . . . . . . . 66
4.10 Thermocouple Schematic . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.1 Timeline of biobutanol history . . . . . . . . . . . . . . . . . . . . . . . . 73
6.1 A simplied life cycle owchart for corn-derived butanol and ethanol . . . 85
6.2 WTP fossil energy use . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.3 WTP petroleum energy use . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.4 WTP GHG emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.5 WTW fossil energy use . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.6 WTW GHG emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
xiii
7.1 Schematic of the laminar ame speed measurement setup . . . . . . . . . 105
7.2 Comparison of the experimental and predicted concentration proles ob-
tained from the oxidation of n-butanol in a JSR at o=1, P=1013 kPa,
t=0.7 s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
7.3 Comparison of the experimental and predicted concentration proles ob-
tained from the oxidation of n-butanol in a JSR at o=1, P=101.3 kPa,
t=0.07 s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
7.4 Comparison of the experimental and predicted concentration proles ob-
tained from the oxidation of n-butanol in a JSR at o=2, P=101.3 kPa,
t=0.07 s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.5 Comparison of the experimental and predicted concentration proles ob-
tained from the oxidation of n-butanol in a JSR at o=0.5, P=101.3 kPa,
t=0.07 s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.6 Comparison of the experimental and predicted concentration proles ob-
tained from the oxidation of n-butanol in a JSR at o=0.25, P=101.3 kPa,
t=0.07 s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
7.7 Reaction pathway diagram for n-butanol oxidation in the JSR at o=0.5,
o=1, o=2, P=101.3 kPa, t=0.07s, and T=1160 K. . . . . . . . . . . . . 119
7.8 Sensitivity of n-butanol concentraion to select reactions in the JSR at
o=1.0, P=101.3 kPa, t=0.07 s, and T=1160 K. . . . . . . . . . . . . . . 120
7.9 Experimental and computed proles obtained from the oxidation of n-
butanol in an atmospheric opposed-ow ame (5.89% C
4
H
9
OH, 42% O
2
). 122
7.10 Reaction pathway diagram for n-butanol oxidation in the opposed-ow
diusion ame at T=858 K, T=1170 K, and T=1520 K . . . . . . . . . . 126
7.11 Sensitivity of n-butanol concentration to select reactions in the atmo-
spheric opposed-ow diusion ame (6% C
4
H
9
OH, 42% O
2
). . . . . . . . 127
7.12 Laminar burning velocities of n-butanol/air mixtures, T=350 K, P=90 kPa.128
8.1 Transesterication Reaction . . . . . . . . . . . . . . . . . . . . . . . . . 143
9.1 Biodiesel FAME and their surrogates . . . . . . . . . . . . . . . . . . . . 154
9.2 Computed proles obtained from the oxidation of methyl decanoate and
methyl butanoate in a JSR at o=1.0, P=1013.25 kPa, t=1 s, 0.1% fuel
mole fraction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
xiv
9.3 One combustion pathway of methyl butanoate that depicts the fate of the
ester moiety. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
9.4 A comparison of the combustion pathways for methyl trans-2-butenoate
(above) and methyl butanoate (below) which lead to unsaturated hydro-
carbons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
9.5 Decomposition of the MD4J radical to 1-octene (C
8
H
16
) and the ME2J
radical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
9.6 Decomposition of the MD2J to methyl 2-propenoate (MP2D) and the
C
7
H
15
radical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
9.7 Abstraction of H atoms from the alpha carbon atom by a reactive radical
species (R) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
9.8 Abstraction of H atoms from the methoxy carbon atom by a reactive
radical species (R) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
9.9 Unimolecular decomposition of MD via scission of C-C bonds in and
around the carbonyl group . . . . . . . . . . . . . . . . . . . . . . . . . . 168
9.10 Experimental (symbols) and computed (lines with symbols) proles ob-
tained from the oxidation of methyl decanoate, n-decane, and methyl bu-
tanoate in a JSR at o=1.0, P=1013.25 kPa, t=1 s, 0.1% fuel mole fraction.170
9.11 Comparison of MD mechanisms (lines with symbols) and experimental
data (symbols) for RME in a JSR at o=1.0, P=101.325 kPa, t=1.0 s [26]. 171
9.12 Critical pressure (P
c
) and critical temperature (T
c
) for C
2
-C
10
methyl esters.173
9.13 Experimental and computed proles obtained from the oxidation of MD
in an atmospheric opposed-ow ame (1.8% MD, 42% O
2
). . . . . . . . . 177
9.14 Reaction pathway diagram for consumption of the MDMJ radical in the
opposed-ow diusion ame at T=1040 K. . . . . . . . . . . . . . . . . . 178
9.15 Reaction pathway diagram for consumption of the MD2J radical in the
opposed-ow diusion ame at T=1040 K. . . . . . . . . . . . . . . . . . 179
9.16 Reaction pathway diagram for consumption of the MD4J radical in the
opposed-ow diusion ame at T=1040 K. . . . . . . . . . . . . . . . . . 180
9.17 Reaction pathways for the formation and consumption of methyl 2-propenoate
in the opposed-ow diusion ame at T=1040 K. . . . . . . . . . . . . . 183
9.18 Reaction pathways leading to the formation of the methoxycarbonyl radi-
cal in the opposed-ow diusion ame at T=1030 K given the experimental
and modeling conditions of Gail et al. [2]. . . . . . . . . . . . . . . . . . 184
xv
9.19 Comparison of proposed MD skeletal mechanism and experimental data
for RME in a JSR at o=1.0, P=101.325 kPa, t=1.0 s [26]. . . . . . . . . 186
9.20 Comparison of proposed MD skeletal mechanism (lines with symbols) and
experimental data (symbols) for RME in a JSR at o=1.0, P=101.325 kPa,
t=1.0 s [26]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
xvi
Acronyms
ABE acetone-butanol-ethanol
ANL Argonne National Laboratory
BDE bond dissociation energy
BP British Petroleum
Bu85 85% butanol-15% gasoline
CG conventional gasoline
CI compression-ignition
DDGS dried grains with solubles
DFT density function theory
DRG directed relation graph
E85 85% ethanol-15% gasoline
EEI Environmental Energy Inc.
FAME fatty acid methyl esters
FFV exible fuel vehicle
FID ame ionization detector
FTIR fourier transform infrared spectroscopy
FU functional unit
GC/FID gas chromatograph ame ionization detector
GC Gas chromatography
GDP gross domestic product
GHG greenhouse gases
GREET Greenhouse Gases, Regulated Emissions, and Energy Use in Transportation
xvii
GSV gas sampling valve
GUI graphical user interface
HACA hydrogen abstraction carbon addition
ICEs internal combustion engines
ID inner diameter
JSR jet stirred reactor
LCA life cycle assessment
LDV light duty vehicle
LHV lower heating value
LOD limit of detection
MB methyl butanoate
MC methyl trans-2-butenoate
MD methyl decanoate
NDIR non-dispersive infrared
NO
a
oxides of nitrogen
NTC negative temperature coecient
OD outer diameter
P
c
critical pressure
PM particulate matter
PTFE polytetrauoroethylene
PTW pump-to-wheel
RME rape seed oil methyl ester
RVP reid vapour pressure equivalent
xviii
SI spark-ignition
T
o
boiling point
T
c
critical temperature
TCD thermal conductivity detector
THC total unburnt hydrocarbons
VOC volatile organic compounds
WTP well-to-pump
WTW well-to-wheel
xix
Statement of Co-Authorship and Copyright
The material presented in this thesis benets from collaborations with a number of re-
searchers. In addition, much of the material has been published in academic journals.
This section details the authors contribution to each section and the right to reproduce
published material without infringing on any copyrights.
All journal articles published from material in this thesis fall under Elseviers copyright
policy
1
. The policy states that the author has the right to include the article in full or
in part in a thesis without obtaining specic permission from Elsevier.
Butanol LCA Studies
The biobutanol life cycle assessment research presented in Chapter 6 was performed in
collaboration with Professor Murray Thomson, Professor Heather MacLean, Professor
Mike Grin, Yimin Zhang, and Sylvia Sleep. S.M. Sarathy conceptualized the research
and was the primary contributer towards literature survey, developing modeling assump-
tions, life cycle assessment modeling, and writing the manuscripts. Yimin Zhang also
contributed signicantly towards literature survey, developing modeling assumptions, life
cycle assessment modeling, and manuscript revisions. Professor Heather MacLean and
Professor Mike Grin contributed towards developing modeling assumptions, research
supervision, and manuscript revisions. Sylvia Sleep contributed towards literature sur-
vey. Professor Murray Thomson provided nancial support to S.M. Sarathy. The above
co-authors have approved the material published in this dissertation.
This study was presented at the following peer reviewed conference proceedings:
1. S.M. Sarathy, Y. Zhang, W.M. Grin, M.J. Thomson, and H. Maclean, Life Cy-
cle Assessment of Biobutanol for use in Transportation Applications. 8th World
Congress of Chemical Engineering Conference, Montreal, Canada, 2009.
Butanol Combustion Studies
The butanol combustion research presented in Chapter 7 was performed in collaboration
with Professor Murray Thomson, Professor Philippe Dagaut, Dr. C. Togbe, Professor
Christine Rouselle, and Professor Fabien Halter. S.M. Sarathys contribution included
1
Available online at http://www.elsevier.com/wps/nd/authorsview.authors/copyright
xx
literature survey, developing chemical kinetic mechanisms, simulating the jet stirred re-
actor and opposed-ow diusion ame, acquiring experimental data in the opposed-ow
diusion ame, and preparing the manuscripts. Professor Murray Thomson contributed
towards nancial support to S.M. Sarathy, research supervision, and manuscript revi-
sions. Professor Philippe Dagauts contribution was research supervision, developing
chemical kinetic mechanisms, manuscript revisions, and nancial support to C. Togbe
who obtained the jet stirred reactor experimental data. Professor Christine Rouselle and
Professor Fabien Halter contributed the laminar ame speed experimental data and sim-
ulations. The above co-authors have approved the material published in this dissertation.
This study was published in the following journals:
1. S.M. Sarathy, M.J. Thomson, C. Togbe, P. Dagaut, F. Halter, C. Mouna

im-
Rousselle. An experimental and kinetic modeling study of n-butanol combustion.
Combustion and Flame, 2009, Vol. 156, 852-864.
2. P. Dagaut, S.M. Sarathy, M.J. Thomson. A Chemical Kinetic Study of n-Butanol
Oxidation at Elevated Pressure in a Jet Stirred Reactor. Proceedings of the Com-
bustion Institute, 2009, Vol. 32, 229-237.
Biodiesel Combustion Studies
The biodiesel combustion research presented in Chapter 9 was performed in collabora-
tion with Professor Murray Thomson, Professor Tianfeng Lu, and Doctor William Pitz.
S.M. Sarathys contribution included literature survey, developing the modied detailed
chemical kinetic mechanism, performing computer simulations, acquiring experimental
data in the opposed-ow diusion ame, and preparing the manuscripts. Professor Mur-
ray Thomson contributed towards nancial support to S.M. Sarathy, research supervi-
sion, and manuscript revisions. Dr. William Pitzs contribution was the original methyl
decanoate chemical kinetic mechanism and additional research supervision. Professor
Tianfeng Lu contributed the algorithm used to reduce the modied detailed chemical
kinetic mechanism developed by S.M. Sarathy. The above co-authors have approved the
material published in this dissertation.
This study has been submitted for publication in the following peer-reviewed confer-
ence proceedings:
xxi
1. S.M. Sarathy, M.J. Thomson, T. Lu, W.J. Pitz. An experimental and kinetic
modeling study of methyl decanoate combustion. Proceedings of the Thirty Third
International Combustion Symposium, 2010, Beijing, China.
2. S.M. Sarathy, M.J. Thomson. Chemical Kinetic Modeling of Biodiesel Combustion.
8th World Congress of Chemical Engineering Conference, Montreal, Canada, 2009.
xxii
Part I
Background and Methods
1
Chapter 1
Introduction
The National Resources Canada 2005 report on energy eciency trends in Canada [1]
indicates that the transportation sector accounts for 30% of total energy use, second only
to the industrial sector. However, transportation produces the largest share of greenhouse
gases (GHG) because the fuels used in transportation are the most GHG intensive. In
addition, Canada spent nearly $61 billion on transportation fuels, the most of any sector.
The transportation sector includes road, air, rail, and marine vehicles; however, the main
source of energy use and GHG was road vehicles used for moving passengers and freight.
From 1990-2005, transportation energy use and GHG increased by more than 30 % due
to an increase in passenger kilometers driven, the consumer shift from cars to minivans
and light trucks, and the increased use of energy intensive modes for transportation.
Although these statistics are for the Canadian economy, the trends are consistent with
other high gross domestic product (GDP) economies.
Road vehicles are powered by internal combustion engines (ICEs) fueled by either
gasoline or diesel. These petroleum derivatives are inherently expensive, and upon
combustion they release large amounts of GHG and other pollutants. Alternatives to
petroleum derived transportation fuels are attractive due to the increasing demand and
limited supply of conventional fossil fuels. Liquid fuels derived from biomass feedstock
(i.e., biofuels) are attracting interest as transportation fuels because they are renewable,
can be locally produced, are more biodegradable, and may reduce net GHG [2]. The
primary driver for using biofuels in the transportation sector is to displace fossil fuel
use. Reducing harmful emissions to the atmosphere is also imperative for mitigating
global warming and sustaining healthy metropolitan areas for human inhabitation. In
addition, the ability for citizens to locally grow and produce their own fuel minimizes the
2
Chapter 1. Introduction 3
dependence on nonrenewable and foreign energy sources.
Currently, gasoline fuel is displaced with bioethanol, while diesel fuel is displaced
with biodiesel. Bioethanol is an alcohol conventionally produced via fermentation of
agriculturally derived starches and sugars. It is blended with gasoline for use in spark-
ignition (SI) engines with only minor modications required to the engine and fueling
systems. Biodiesel is dened as a mixture of mono-alkyl esters of long chain fatty acids
derived from vegetable oils or animal fats [3]. Biodiesel can be used in its pure form
or it can be blended with petroleum diesel without major modications to the existing
compression-ignition (CI) engine and fuel distribution infrastructure. In 2007, biofuels
accounted for over 1.5% of global transport fuels [4] with bioethanol and biodiesel con-
tributing an estimated 47 billion liters and 8 billion liters, respectively [5]. The global
demand for biofuels has tripled since 2000, and strong growth is expected in the near
future due to favourable policies from North American and European governments.
A recent review on biofuels provides a unique perspective on the environmental and
societal impacts of biofuels [6]. The rapid policy-driven growth of biofuel use has led to
serious environmental and food security concerns. Current biofuel technologies compete
with the food industry for feedstock, and the diversion of corn, rice, and oilseeds to biofuel
production is cited as the cause of rising food prices and global food shortages. In addi-
tion, large amounts of forest land are being destroyed for biofuel feedstock production,
leading to a loss in biodiversity and carbon-rich sinks. The competition for fresh water
resources presents an additional barrier towards widespread biofuel use. Despite these
challenges, advances in biofuel feedstock and production technologies can ameliorate the
negative impacts of biofuels. Along with environmental stewardship, energy conserva-
tion, eciency improvements, and other renewable energy technologies (e.g., solar, wind,
geothermal, etc.), biofuels can safely be part of a diverse energy portfolio that reduces
fossil fuel consumption.
1.1 Research Motivation
The rapid increase in biofuel use has sparked an equally rapid growth in research on
biofuel sustainability assessment and combustion properties. Biofuels can either benet
or harm the environment, so sustainability assessment research attempts to determine the
net environmental impacts associated with biofuel production and use. The most widely
used tool for sustainability assessment is the life cycle assessment (LCA) methodology,
Chapter 1. Introduction 4
which determines the biofuels environmental performance based on a set of user-dened
metrics, such as fossil energy input and GHG output. Biofuel combustion research studies
the fundamental combustion properties in order to improve vehicle performance and
minimize harmful emissions. One important combustion tool is the chemical kinetic
mechanism, which describes the molecular level transformation of reactants (i.e., fuel
and air) into products via a series of elementary steps. These mechanisms can be used to
predict ignition properties, heat release rates, amounts of emissions, and the types and
levels of intermediate species [7] in any combustion system.
The broad research questions that this dissertation addresses are:
Do biofuels oer better environmental performance than fossil fuels?
How does biofuel combustion dier from fossil fuel combustion?
LCAs and chemical kinetic mechanisms of bioethanol fuel are already under intense
research, so such work is beyond the scope of this dissertation. Recently, biobutanol has
attracted attention with British Petroleum and DuPont announcing they would begin
selling sugar beet derived butanol as a gasoline blending component in the United King-
dom [8]. This announcement, in combination with reported research and development
advances in biobutanol production, and cited fuel property advantages of biobutanol
compared to bioethanol, suggest that sustainability assessment and combustion research
on biobutanol is warranted.
Biodiesel sustainability assessment research is already well established, but funda-
mental combustion research is limited due to biodiesels complex composition. Biodiesel
is typically comprised of a mixture of saturated and unsaturated fatty acid alkyl esters
(i.e., fatty acid methyl esters (FAME)
1
) with chain lengths ranging from 12 to 18 carbon
atoms. Developing chemical kinetic models for biodiesel has been challenging due to the
large size of the fatty acid alkyl esters found in practical fuels. The added complexity
of varying chain length and degrees of unsaturation has led to the use of surrogate fuels
of well characterized composition for chemical kinetic modeling. Well validated chemical
kinetic mechanisms exist for short chain FAME surrogates; however, there are few vali-
dated chemical kinetic mechanisms for long chain FAME surrogates, so this dissertation
lls the void.
1
Biodiesel can also be comprised of ethyl esters, but this study is only concerned with methyl esters
since these constitute the majority of biodiesel in production today.
Chapter 1. Introduction 5
1.2 Dissertation Objectives and Layout
The primary focus of this dissertation is to research the combustion kinetics of biobutanol
and FAME. The study proceeds by performing fundamental combustion experiments and
then using the experimental data to validate chemical kinetic mechanisms for the biofuels.
In addition, since biobutanol is a new biofuel that has not been critically assessed for
sustainability, this dissertation also performs an LCA of biobutanol.
This dissertation is divided in three parts. Part I includes this introduction, as well
as background material relevant to both biobutanol and biodiesel combustion chemistry.
Part II contains all research material related to biobutanol, which includes an LCA of
biobutanol and an experimentally validated chemical kinetic mechanism. Part III is
dedicated to research on biodiesel, which focuses on creating an experimentally validated
chemical kinetic mechanism for long chain FAME. Below is a specic list of objectives
for each part of this dissertation.
Part I Background and Methods
Chapter 1 Present the research motivation and objectives of this dissertation
Chapter 2 Provide a background on combustion chemistry in practical applications
Chapter 3 Discuss the modeling of combustion chemistry
Chapter 4 Describe the experimental methods used for validating chemical kinetic
mechanisms
Part II Biobutanol
Chapter 5 Provide background information related to biobutanol
Chapter 6 Present the LCA of biobutanol
Chapter 7 Present the validated chemical kinetic mechanism for biobutanol combustion
Part III Biodiesel
Chapter 8 Provide background information related biodiesel
Chapter 9 Present the validated chemical kinetic mechanism for biodiesel combustion
Chapter 1. Introduction 6
Part IV Closing
Chapter 10 Summarizes the contributions of this dissertation
Literature Cited
[1] NRCAN, Energy eciency trends in canada, 1990 to 2005, Natural Resources
Canada Oce of Energy Eciency, Tech. Rep., 2009.
[2] A. Demirbas, Importance of biodiesel as transportation fuel, Energy Policy, vol. 35,
no. 9, pp. 46614670, September 2007.
[3] ASTM, ASTM D 6751 specication for biodiesel fuel blend stock (B100) for
middle distillate fuels, in ASTM Book of Standards. ASTM, 2003.
[4] R. Sims, M. Taylor, J. Saddler, and W. Mabee, From 1st to 2nd generation biofuel
technologies - an overview of current industry and RD&D activities, International
Energy Agency and IEA Bioenergy, Tech. Rep., 2008.
[5] REN21, Renewables 2007 global status report, Renewable Energy Policy Network
for the 21st Century, Tech. Rep., 2008.
[6] L. P. Koh and J. Ghazoul, Biofuels, biodiversity, and people: Understanding the
conicts and nding opportunities, Biological Conservation, vol. 141, no. 10, pp.
24502460, 2008.
[7] C. K. Westbrook, W. J. Pitz, P. R. Westmoreland, F. L. Dryer, M. Chaos, P. Osswald,
K. Kohse-Hoeinghaus, T. A. Cool, J. Wang, B. Yang, N. Hansen, and T. Kasper, A
detailed chemical kinetic reaction mechanism for oxidation of four small alkyl esters
in laminar premixed ames, Proceedings of the Combustion Institute, vol. 32, no.
Part 1, pp. 221228, 2009.
[8] G. Hess, BP and DuPont to make biobutanol, Chemical & Engineering News,
vol. 84, pp. 910, 2006.
7
Chapter 2
Background
Since the late 1800s when Otto invented the SI engine and Diesel invented the CI engine,
ICEs have become a leading source of stationary and mobile power. ICEs convert the
fuels chemical energy into mechanical energy via oxidation (i.e., combustion) within the
engine. Therefore, the reactant fuel-air mixture and the combustion products are the
working uids of the engine. The dierence between the SI and CI engine lies in the
method by which the fuel-air mixture is introduced to the combustion chamber and ig-
nited. The method of ignition determines the key engine characteristics, including the fuel
requirements, operating temperatures and pressures, emission formation mechanisms,
and performance and eciency [1]. Table 2.1 summarizes several unique characteristics
of SI and CI engines.
The vast majority of ICEs in mobile power applications burn petroleum derived liquid
hydrocarbon fuels. However, early pioneers in the auto industry envisioned the use of
liquid fuels derived from biomass; Henry Fords Model T was designed to run on ethanol
while Rudolf Diesel operated his CI engine on peanut oil. This chapter begins with a
brief background of SI and CI engines, fuel properties, and combustion emissions, so that
the reader understands the importance of studying the combustion chemistry of biofuels
1
.
The following chapters then focuses on combustion kinetics, including its importance, the
development of kinetic mechanisms, and the use of computer simulations as a modeling
tool.
1
A thorough description of engine fundamentals is available in Heywoods text Internal Combustion
Engine Fundamentals [1].
8
Chapter 2. Background 9
Table 2.1: Selected characteristics of SI and CI engines
SI CI
Ignition Mode spark ignition compression ignition
Combustion Mode premixed nonpremixed
turbulent ame turbulent diusion ame
Petroleum Fuel gasoline diesel
Biofuel alcohols fatty acid alkyl esters
Compression Ratio 8 to 12 12 to 24
2.1 Reciprocating ICEs
In reciprocating ICEs, the piston moves cyclically up and down in the cylinder chamber
to produce work. The ratio of the maximum cylinder volume (i.e., when the piston is
at the bottom of its stroke) to the minimum cylinder volume is called the compression
ratio. Typically, to generate one power stroke, the piston goes through a four-stroke cycle
which consists of the following [1]:
1. The intake stroke draws fresh mixture into the cylinder chamber by opening the
intake valve and moving the piston from the top of the cylinder to the bottom.
2. During the compression stroke the intake valve closes and the piston moves back
towards the top of the cylinder, compressing the cylinder mixture. Combustion
begins near the end of the compression stroke causing a rapid pressure rise.
3. The expansion stroke occurs as the rapid pressure rise in the cylinder forces the
piston downwards to the bottom of the cylinder chamber. The work generated dur-
ing expansion, which is ve times greater than the work used during compression,
turns a crank shaft that delivers power to the vehicles wheels.
4. Finally, the exhaust stroke pushes the burned gases out of the cylinder by moving
the piston upwards and opening the exhaust valve. When the piston reaches the
top of the cylinder chamber, the exhaust valve closes and the intake valve opens,
and the four-stroke cycle is reciprocated.
Chapter 2. Background 10
2.1.1 SI Engines
SI engines are characterized by air and fuel being premixed prior to entering the cylinder
chamber through the intake valve. Air-to-fuel ratios in SI engines are typically near
stoichiometric. As the piston moves upwards and the cylinder volume decreases, the
premixed cylinder gas is compressed to 0.8-1.4 MPa (8-14 atm). A typical compression
ratio in SI engines is 8 to 12, which results in less work per stroke when compared to CI
engines [1]. Lower compression ratios are required in SI engines to minimize auto-ignition
(i.e., knocking) of the air-fuel mixture during the compression stroke. Near the top of the
compression stroke, a spark plug ignites the cylinder gases and propagates a turbulent
ame through the cylinder chamber. The rapid pressure and temperature rise forces the
piston downwards through the expansion stroke. Finally, the exhaust stroke forces the
burned gases out of the cylinder and the process is repeated.
2.1.2 CI Engines
In CI engines, only air enters the cylinder during the intake stroke. The cylinder air is
then compressed to approximately 4 MPa (40 atm) and 800 K. A typical compression
ratio in CI engines is 12 to 24, allowing for a greater amount of work to be done during
each cycle [1]. Near the end of the compression stroke, fuel is injected directly into the
cylinder chamber. The liquid fuel jet impinges upon the hot cylinder air and begins to
vaporize. Small pockets of premixed fuel and air then auto-ignite creating additional
heat and radicals sucient to generate a diusion ame which consumes the remaining
liquid fuel jet. The exhaust valve opens near the end of the expansion stroke and the
burned gases are exhausted as the piston moves back upwards to its starting position.
2.2 Fuel Properties
The nature of the ignition process in SI and CI engines determines each engines fuel
requirements. In SI engines, ignition is initiated by a spark and it is important for the
fuel-air mixture to avoid autoignition. However in CI engines, fuel autoignition is desired
in order to initiate the combustion process during the expansion stroke. ASTM standards
for petroleum derived gasoline [2] and diesel [3] provide specications for fuel properties
such as density, viscosity, volatility, autoignition characteristics, composition, stability,
and seasonal performance. It is important for biofuels to have properties similar to their
Chapter 2. Background 11
hydrocarbon counterparts, so that major changes are not required to the fueling system
and engine.
A fuels ability to resist autoignition in SI engines is known as its antiknocking ten-
dency, and it is quantied by its octane number. The octane number is based on the
fuels knocking tendency relative to n-heptane (octane number 0) and iso-octane (2,2,4-
trimethylpentane, octane number 100) [1]. Higher octane numbers indicate a greater
resistance to knock, and are therefore desirable. The octane number of a hydrocarbon
fuel decreases with increasing chain length. Branched hydrocarbons have higher octane
numbers than straight hydrocarbons of the same carbon number because the length of
the basic chain is reduced. SI engine fuels (i.e., gasoline) have octane ratings ranging
from 87 to 105, and mainly consist of straight and branched hydrocarbons of less than
12 carbon atoms with a smaller amount of aromatic species. The addition of oxygenated
compounds, such as ethanol, to hydrocarbon fuel increases the octane number because
oxygenates have greater antiknocking tendencies.
A fuels ability to autoignite in CI engines is measured by its cetane number, and it
is inversely related to octane number. The cetane number is based on the fuels ability
to autoignite relative to n-hexadecane (cetane number 100) and 1-methylnaphthalene
(cetane number 0). For hydrocarbon fuels, the cetane number increases with increasing
chain length and decreases with branching and cyclication. CI engine fuels (i.e., diesel)
have cetane ratings ranging from 40 to 55 [1], and consist of straight chain hydrocarbons
between 10 and 20 carbon atoms in length with lesser amounts of branched and aromatic
hydrocarbons. Alcohols have low cetane numbers, thereby making them dicult to use
in CI engines. Fatty acid methyl ester species like those found in biodiesel have high
cetane numbers, making them suitable for CI engines.
Besides octane/cetane rating, another important fuel property in SI and CI engines
is the amount of energy per unit of volume. Fuels are sold on a volumetric basis, but
it is energy that powers a vehicle. The energy and volume of a fuel are related through
the density (i.e., kg/L) and the lower heating value (LHV) (i.e., MJ/kg). The LHV is
dened as the amount of energy (i.e., MJ) released during the combustion of a specied
mass of fuel (i.e., kg) at 25

C and returning the combustion products to 25

C, and then
subtracting the latent heat of vaporization of the water vapor formed during combustion.
Thus, the LHV assumes that water in the combustion products is in the vapor state,
and energy is not recovered by condensing it out of the combustion gas. The volumetric
heating value (i.e., MJ/L) is obtained by multiplying the fuels LHV by the fuels density,
Chapter 2. Background 12
and it is this value that is important for consumers. Fuels with a larger volumetric heating
value increase fuel economy (i.e., L/100 km) because the amount of energy per unit of
fuel volume is greater. Biofuels tend to have smaller volumetric heating values than their
hydrocarbon counterparts; therefore, it is expected that the fuel economy of a vehicle
powered by biofuel is lower than when powered by petroleum. It should be noted that
some studies have reported an improvement in engine eciency when using biofuels,
which osets the lower volumetric heating value, and results in no net change in fuel
economy [4].
2.3 Combustion Emissions
ICEs are a major source of air pollutants and GHG emissions. These emissions have
short-term and long-term health eects on humans, which include irritation of the eyes
and respiratory tract, severe respiratory illnesses, heart disorders, and cancer. Environ-
mental eects include global warming caused by GHG emissions, ozone layer depletion,
acidication, and urban smog formation [5].
Both SI and CI engines emit oxides of nitrogen (NO
a
), carbon monoxide (CO), par-
ticulate matter (PM), and total unburnt hydrocarbons (THC). These emissions are
regulated by government agencies, so vehicles must meet stringent emission standards.
Combustion in SI engines creates low levels of PM, high levels of CO, THC, and NO
a
,
but the use of three-way catalytic converters greatly reduces the tail-pipe emissions of
these compounds. CI engine combustion generally produces lower levels of THC and
CO, comparable levels of NO
a
, and higher levels of PM. Until recently, tail-pipe emis-
sions of CO and THC were reduced using two-way catalytic converters, and little exhaust
treatment was conducted to reduce NO
a
and PM emissions. However, new environmen-
tal regulations are introducing the use of NO
a
traps and diesel particulate lters for CI
engines.
Typically, biofuels produce similar amounts of regulated emissions as hydrocarbon
fuels, but there is one notable dierence. The use of oxygenated fuels (e.g., biofuels)
has been shown to be an eective way of reducing soot emissions in diesel engines [5, 6].
Oxygenated fuels reduce soot formation by i. sequestering carbon atoms from forming
soot by creating carbon-oxygen bonds, and ii. reducing the aromatic content compared
Chapter 2. Background 13
to petroleum fuels
2
. A recent study by Pepiots-Desjardins et al. [7] studied the sooting
tendency of various oxygenated (e.g., alcohols, esters, aldehydes, etc.) and hydrocarbon
compounds, and concluded that oxygenated fuels have soot reducing eciencies that are
directly related to nature of the functional group.
In addition to the regulated emissions mentioned above, there are several unregulated
emissions that may be of concern. Carbon dioxide is a major cause of global warming
and its emission is currently not regulated by government agencies. Oxygenate emissions,
such as aldehydes and ketones, may become more important when using oxygenated fuels
because the fuel bound oxygen could lead to direct formation of oxygenated compounds.
A study by Jacobson [8] concluded that widespread ethanol use may increase the risk of
cancer and ozone-related illness due to higher aldehyde emissions and increased unburnt
ethanol emissions, which break down to acetaldehyde in the atmosphere. It is uncertain
whether or not biodiesel leads to higher oxygenate emissions since engine studies have
shown both increases and decreases when compared to diesel [4]. Combustion chem-
istry studies of biofuels can help determine the importance of oxygenated emissions by
elucidating the role of fuel bound oxygen during combustion.
The composition of the pollutant gases in the cylinder chamber at the end of the ex-
pansion process varies depending on the engine operating parameters. The concentration
of the pollutants can be calculated assuming chemical equilibrium, as described in the
next section, but these values tend to dier greatly than measured values. This discrep-
ancy is because the combustion products cool rapidly during the expansion process, and
the chemical reactions controlling pollutant formation become rate limited (i.e., they
cannot achieve equilibrium). The pollutant concentrations are essentially frozen at
their higher temperature values. Detailed chemical mechanisms and their corresponding
kinetic parameters are required for an accurate calculation of pollutant concentrations
[1]. Therefore, combustion chemistry studies, such as those presented in this thesis, are
required to design engines that curtail pollutant emissions.
2
Aromatic hydrocarbons lead directly to soot formation.
Literature Cited
[1] J. Heywood, Internal Combustion Engine Fundamentals. McGraw-Hill Book Com-
pany, New York., 1988.
[2] ASTM, ASTM D 4814 standard specication for automotive spark-ignition engine
fuel, in ASTM Book of Standards. ASTM, 2003.
[3] ASTM, ASTM D 975 standard specication for diesel fuel oils, in ASTM Book of
Standards. ASTM, 2003.
[4] M. Lapuerta, O. Armas, and J. Rodriguez-Fernandez, Eect of biodiesel fuels on
diesel engine emissions, Progress in Energy and Combustion Science, vol. 34, no. 2,
pp. 198223, April 2008.
[5] A. Agarwal, Biofuels (alcohols and biodiesel) applications as fuels for internal com-
bustion engines, Progress in Energy and Combustion Science, vol. 33, no. 3, pp.
233271, June 2007.
[6] J. Song, K. Cheenkachorn, J. Wang, J. Perez, A. L. Boehman, P. J. Young, and F. J.
Waller, Eect of oxygenated fuel on combustion and emissions in a light-duty turbo
diesel engine, Energy & Fuels, vol. 16, no. 2, pp. 294 301, 2002.
[7] P. Pepiot-Desjardins, H. Pitsch, R. Malhotra, S. Kirby, and A. Boehman, Structural
group analysis for soot reduction tendency of oxygenated fuels, Combustion and
Flame, vol. 154, pp. 191208, 2008.
[8] M. Z. Jacobson, Eects of ethanol E85 versus gasoline vehicles on cancer and mor-
tality in the united states, Environmental Science and Technology, vol. 41, no. 11,
pp. 41504157, June 2007.
14
Chapter 3
Modeling Combustion Chemistry
Combustion in ICEs is a complex process involving fuel atomization, vaporization, fuel-air
mixing, ignition, and combustion. For example, in a CI engine liquid fuel is injected as a
high velocity spray into the combustion chamber, where it vaporizes upon impingement
with high-temperature high-pressure cylinder gases. Low temperature reactions then
spontaneously ignite portions of premixed fuel and air causing rapid heat release. The
remaining fuel spray is then consumed in a high temperature diusion ame, and burned
gases are produced through the entire expansion process. This unsteady, heterogeneous,
3-dimensional process is challenging to model, and it is dicult to decouple mixing
processes from chemical kinetic processes [1].
Computer simulations based on the KIVA code [2] are capable of combining uid
dynamics, spray dynamics, chemically reacting ows, and heat and mass transfer in
an engine cylinder to predict ignition behavior, pollutant formation, energy release, and
other features of engine operation. Such codes are widely used in the automobile industry
to increase fuel economy and reduce emissions. Typically, these engine simulations are
computationally expensive, so simplications to uid dynamics, spray dynamics, and
elementary chemical kinetics are required. However, reducing the chemical kinetic model
(i.e., mechanism) reduces chemical delity and limits our ability to fully understand
combustion chemistry. This chapter describes how detailed chemistry is modeled in
idealized combustion systems, so that the eects of molecular structure on combustion
and emissions can be understood.
15
Chapter 3. Modeling Combustion Chemistry 16
3.1 Chemical Kinetics
Many processes in the engine, including reactions in the ame zone which determine heat
release, reactions controlling ignition, and air pollutant formation mechanisms, occur
at times when temperature and pressure are changing rapidly. These nonequilibrium
processes depend on the rate of each individual chemical reaction (i.e., reaction kinetics),
which are governed by the temperature and the concentration of reactants.
The rates at which reactant species are consumed and product species are produced in
a kinetically controlled process is governed by the law of mass action. For the elementary
reaction in Equation 3.1, the law of mass action states that the rate at which reactants
are consumed is proportional to the product of concentration of each reactant raised to its
stoichiometric coecient, as shown in Equation 3.2. The forward reaction rate constant
(/
)
) shown in the equation follows the Arrhenius form and is further discussed in a later
section.
c +/1 = cC +d1 (3.1)

d
dt
= /
)
[]
o
[1]
o
(3.2)
A comprehensive list of chemical reactions and their rates (i.e., a chemical kinetic
mechanism) is required to accurately predict the rate of energy release, soot and pollu-
tant formation, ignition behaviour, knocking limits, and cool ame characteristics [3, 4, 5].
Rather than attempting to validate the detailed kinetic model in an engine, a better op-
tion is to study combustion chemistry and ame structure in idealized chemically react-
ing ow systems (e.g., an opposed-ow diusion ame) [6]. The combustion phenomenon
observed in the laboratory experiment can then be used to validate a chemical kinetic
mechanism and understand combustion performance in an engine. Furthermore, com-
prehensive chemical kinetic mechanisms validated against a wide range of experimental
data provide the foundation for the reduced mechanisms used in engine simulations [7].
3.2 Computer Simulations for Mechanism Validation
Chemical kinetic mechanisms describe the molecular level transformation of reactants
(i.e., fuel and air) into products via a series of elementary steps. The mechanism valida-
tion rst requires a model describing the geometry and operating regime of the specic
Chapter 3. Modeling Combustion Chemistry 17
combustion application. A large number of dierential equations describing the mass,
momentum, energy, and species concentration are numerically integrated to generate
concentration proles for reactants, intermediates, and products [8]. The computed pro-
les are then validated against experimental data from one or more well-characterized
combustion apparatuses.
A chemical kinetic mechanism is typically validated against an idealized chemically
reacting ow system. The experimental setups modeled in the present study included
opposed-ow diusion ames, jet stirred reactors, and premixed laminar ames (i.e.,
laminar ame speed). This section describes the governing equations used for modeling
chemically reacting ow systems. Numerical modeling is not the focus of this dissertation
study, so a complete derivation of equations for the various combustion systems (e.g.,
opposed-ow diusion ames, jet stirred reactors, and premixed laminar ames) and their
solution methodology is not presented herein. The reader is directed to the CHEMKIN
Theory Manual for further elaboration on the specic computer codes used for numerical
modeling [9].
3.2.1 Governing Equations for Chemically Reacting Flows
Chemically reacting ow problems are mathematically formulated using equations for
conservation of mass, momentum, energy, and concentration of chemical species, along
with thermodynamic relationships [5, 6, 10]. The chemical kinetic mechanism couples
chemical species concentrations with the energy equation via the enthalpy of reaction.
A set of ordinary dierential equations for species and energy, with time as the indepen-
dent variable, make up the conservations equations for problems where spatial transport
is negligible (e.g., plug ow reactors, perfectly stirred reactors, etc.). When transport pro-
cesses are important (e.g., laminar ames), the conservation equations becomes a partial
dierential equations, with time and space as the independent variables. The compu-
tational cost for kinetically controlled problems is small, but when transport processes
are included the computational load increases dramatically [5]. The following sections
discuss the governing equations for modeling chemically reacting ow systems comprised
of laminar gaseous ows.
Chapter 3. Modeling Combustion Chemistry 18
Conservation of Mass
Mass is always conserved in a system, and therefore, in a steady state process the rate at
which mass enters a dierential element (i.e., a point in space) is equal to the rate at which
it leaves the element. In uid mechanics, the conservation of mass is mathematically
formulated using the continuity equation shown in Equation 3.3.
j
t
+ (j ) = 0 (3.3)
where
j is the uid density
t is the time
is the uid velocity vector
() is the divergence operator
The equation indicates that the rate of change in mass with respect to time in a
dierential element,
j
|
, plus the net mass ow into and out of that element, (jV), is
zero.
Conservation of Momentum
The conservation of momentum (i.e., Navier-Stokes equations) together with the conti-
nuity equation for conservation of mass are the fundamental formulations in uid me-
chanics. The general dierential form of the Navier-Stokes equations where gravity is the
only acting force is shown in Equation 3.4.
j
t
+
j = )
-&)occ
+ )
oco
=

t j +j p (3.4)
where
j is the uid density
t is the time
is the uid velocity vector
j is the pressure representing a surface force, f
-&)occ

t is the viscous stress tensor representing a surface force, f


-&)occ
p is the gravitational force constant representing a body force, f
oco
Chapter 3. Modeling Combustion Chemistry 19
The equation states that the change in momentum with respect to time in a dierential
element, j

|
, plus the contribution of convection on momentum, j

\ , is equal to the
contribution of viscous stress on momentum,

t, minus the contribution of pressure on


momentum, j, plus the contribution of gravitational forces on momentum, j p.
Conservation of Species
The continuity equation presented above denes mass conservation in a uid ow, but it
does not provide any distinction on the chemical species present in the ow. However, the
mass conservation of individual species is important in chemically reacting ow systems
consisting of a multicomponent gaseous mixture. The mass fraction of an individual
species is shown in Equation 3.5.
1
I
=
j
I
j
(3.5)
where
1
I
is the mass fraction of the /
|
species, and
1

I=1
1
I
= 1
j is the total uid density
j
I
is mass density of the /
|
species, and
1

I=1
j
I
= j
The chemical composition of a gaseous mixtures in a dierential element can be
derived from species mass conservation equations. The mass conservation of a the /
|
species in an element is altered by homogeneous chemical reactions, molecular diusion,
and convection, as shown in Equations 3.6
j
1
I
t
+j 1
I
= .
I
\
I

J
I
(3.6)
where
j is the uid density
t is the time
is the uid velocity vector
1
I
is the mass fraction of the /
|
species

J
I
is the diusive mass ux vector of the /
|
species
.
I
is the net molar production rate of the /
|
species
\
I
is the molecular weight of the /
|
species
Chapter 3. Modeling Combustion Chemistry 20
The equation states that the change in concentration (i.e., mass fraction) of the /
|
species with respect to time in a dierential element, j
Y

|
, plus the contribution of
convection on concentration, j 1
I
, is equal to the contributions of chemical reactions
on concentration, .\
I
, minus the contribution of molecular diusivity on concentration,

J
I
.
The diusivity (i.e., diusion mass ux) of a species, ,
I
, can be described using Ficks
law as shown in Equation 3.7. The theory states that the diusivity depends linearly on
the negative concentration gradient multiplied by the binary diusion coecient, 1
I
.

J
I
= j
1
I
A
I
1
I
A
I
= j
\
I

\
1
I
A
I
(3.7)
where

J
I
is the diusive mass ux vector of the /
|
species
j is the uid density
1
I
is the mass fraction of the /
|
species
A
I
is the mole fraction of the /
|
species
1
I
is the binary diusion coecient
\
I
is the molecular weight of the /
|
species

\ is the mean molecular weight of the mixture


Conservation of Energy
Thermal energy is conserved in chemically reacting ow systems, and the energy equation
is used as the basis for such systems. The energy equation is used to describe the
temperature prole of a chemically reacting ow, which aects processes such as chemical
reaction, convection, and molecular diusion. The thermal energy equation, shown in
Equation 3.8, stems from the rst law of thermodynamics, and assumes ideal gases, low
Mach numbers. and Fouriers law for heat conduction.
jc
j
1
t
+ jc
j

I
1 = (`1) j
1

I=1
c
j,I
1
I

I
1
1

I=1

I
.
I
\
I
+
oo
(3.8)
where
c
j
is the constant pressure heat capacity of the /
|
species
j is the uid density
Chapter 3. Modeling Combustion Chemistry 21
t is time
1 is the temperature
` is the thermal conductivity
1
I
is the mass fraction of the /
|
species

I
is the uid velocity vector of the /
|
species
.
I
is the net molar production rate of the /
|
species

I
is the enthalpy of formation of the /
|
species
\
I
is the molecular weight of the /
|
species

oo
is the radiative heat transfer
The equation states that the change in thermal energy with respect to time in a
dierential element, jc
j
T
|
, plus the thermal energy convected to the element by the
temperature gradient, jc
j
1, is equal to the contribution of thermal heat conduc-
tion (i.e., Fouriers law) on thermal energy, (`1), minus the contribution of thermal
diusivity on thermal energy, j
1

I=1
c
j,I
1
I

I
1, minus the contribution of heat from
chemical reaction on thermal energy,
1

I=1

I
.
I
\
I
, plus the contribution of radiative heat
transfer on the the element,
oo
.
3.3 Solving the Governing Equations
The governing equations can be solved using a numerical solver that evaluates the chemi-
cal kinetic, thermodynamic, and transport properties in each dierential element as time
proceeds. This study models chemically reacting ow systems using the CHEMKIN
software package [11]. CHEMKIN provides modeling of a wide range of combustion ap-
paratuses, including shock tubes, premixed ames, diusion ames, and partially and
perfectly stirred reactors. Chemical kinetic mechanisms are coupled with thermochem-
ical data for all the species in the mechanism to calculate forward and reverse reaction
rates. Transport properties for the species are also included when attempting to model a
combustion process in which transport processes are rate-controlling (e.g., opposed-ow
diusion ames).
The combustion setups were modeled using the CHEMKIN 4.1 software package. The
rst step used the CHEMKIN 4.1 graphical user interface (GUI) to set up a diagram of
the experimental apparatus, including all reactant and product streams. The next step
Chapter 3. Modeling Combustion Chemistry 22
was to generate the linking les for the numerical code. This required using the pre-
processors to access three important information les: i.) the chemical kinetic database;
ii.) the thermodynamic database; and iii.) the transport database. More information
on these les is given below. The CHEMKIN Gas-Phase Interpreter reads the rst
two les and generates the CHEMKIN Linking File. The third le is used by the
TRANSPORT Preprocessor to generate the Transport Linking File when modeling
systems where transport processes are important.
Next, the characteristics of the chemically reacting ow system and inlet ows were
input. This includes the velocity of each inlet stream, initial concentrations, pressure,
physical conguration, temperature, and a number of solution method options. The
model was run, and the numerical simulation output a text le containing the solution.
The Solution Export Utility was used to convert this text le into a comma separated
values le format that was readable by Microsoft Excel.
This section discuss the the three input les required by CHEMKIN for solving the
chemically reaction ow system problem. The development of these input les is the
primary focus of this dissertation, so a thorough background is presented for the reader
to appreciate this studys contributions.
3.3.1 Chemical Kinetic Database
The chemical kinetic database identies all the gaseous species present, and it provides
a user-dened chemical kinetic mechanism for the production and consumption of these
species. The chemical kinetic mechanism details each reaction taking place and the ap-
propriate reaction rate parameters in the modied Arrhenius form, as shown in Equation
3.12. The gas phase kinetic le conforms to the CHEMKIN input format. Additional
information of the development of chemical kinetic mechanisms is provided in Section
3.4.
The chemical source term, .
I
, describes the net molar production rate of the /
|
species, and it appears in Equations 3.6 and 3.8. The chemical kinetic mechanism con-
tains the information needed to evaluate .
I
. The chemical system consisting of ` species
and ` reversible reactions can be expressed as
.

I=1
i

aI
[A
I
]

I=1
i

aI
[A
I
] (3.9)
where
Chapter 3. Modeling Combustion Chemistry 23
i

aI
is the stoichiometric coecient of the /
|
reactant species in the :
|
reaction
i

aI
is the stoichiometric coecient of the /
|
product species in the :
|
reaction
[A
I
] is the molar concentration of the /
|
species
The molar production of the /
|
species, .
I
, is expressed as
.
I
=
A

I=1
(i

aI
i

aI
)c
a
(3.10)
where
c
a
is the progress variable of the :
|
reaction, given in Equation 3.11
c
a
= /
)
a
.

I=1
[A
I
]
i

a
.

I=1
[A
I
]
i

(3.11)
Each :
|
reversible reaction is characterized by a forward reaction rate, /
)
a
following
the Arrhenius form shown in Equation 3.12. The reverse reaction rate constant, /

a
, is
calculated from thermochemistry.
/
)
a
= (1)
a
exp
1
o
1 1
(3.12)
where
/
)
a
is the reaction rate constant
is the pre-exponential collision frequency factor in
cn
3
nc|-
1 is temperature in kelvin
: is the temperature dependence factor
1
o
is the activation energy in
co|
nc|
1 is the ideal gas constant in
co|
nc|1
The reaction rate constant, /
)
a
, in Equation 3.12 depends on the temperature, activa-
tion energy, and the collision frequency factor. A temperature dependence term is incor-
porated into the equation (i.e., : = 0) for reactions that exhibit non-Arrhenius behaviour
over the range of temperatures encountered in combustion. Typically, experiments are
conducted to determine the coecients in the rate equation; however estimations and
calculations using ab initio quantum chemistry methods are also employed. This study
Chapter 3. Modeling Combustion Chemistry 24
uses, wherever possible, rate coecients from published experimental and computational
studies available through the NIST Chemical Kinetics Database [12]. For reactions
that have not been studied previously, this study estimates rate coecients heuristically.
3.3.2 Thermochemical Database
Thermochemical data for each species in the chemical kinetic mechanism are required for
CHEMKIN to calculate thermodynamic properties, thermal transport properties, and
reaction equilibrium constants. Contained within the thermochemical data le are the
species name, elemental composition, electronic charge, and phase. In addition, fourteen
polynomial tting coecients are provided to calculate the constant pressure molar heat
capacity (C
0
j
), molar enthalpy of formation (
)
H
0
), and molar entropy of formation
(S
0
j
) at any temperature. From these calculated properties, CHEMKIN can calculate
other important thermochemical properties, such as the constant volume heat capacity,
internal energy, Gibbs free energy, and Hemholtz free energy. Mass based properties are
generated by dividing the property in molar units by the molecular weight [9].
Both computational and experimental methods can be used to determine thermo-
chemical data properties, and the NIST Chemistry WebBook has a good compilation
of previously published data [13]. For larger molecules which have not been experimen-
tally studied and for which quantum chemical calculations are computationally expensive,
thermochemical properties can be estimated based on group additivity methods. Ben-
son [14] has proposed a systematic way of estimating thermochemical properties for a
molecule from data on the bonded atomic groups which comprise it. The additivity
law determines the property X of a complex molecule by adding the tabulated bond
properties for simple molecules (e.g.,

A
nc|cc&|c
=

A
ocao-
). The THERGAS [15] and
THERM [16] softwares, which is based on Bensons method, are used in the present study
to estimate thermochemical properties of new species. The user inputs the structure of
the molecule in a specied format, and the computer determines its thermochemical
properties by adding together tabulated bond properties.
3.3.3 Transport Database
Combustion is typically a combination of chemical kinetic driven processes (i.e., produc-
tion and destruction of species) and transport driven processes (i.e., convection, diusion,
and conduction). In certain combustion applications, such as a perfectly stirred reactor or
Chapter 3. Modeling Combustion Chemistry 25
plug-ow reactor, the overall rate is assumed to be kinetically controlled since the trans-
port processes occur innitely fast. However in other cases, such as laminar and diusion
ames, the transport processes are rate-controlling. Therefore, the molecular transport
of species, momentum, and energy in the gas mixture must be evaluated from the dif-
fusion coecients, viscosities, thermal conductivities, and thermal diusion coecients.
CHEMKIN determines these temperature and pressure dependent ow properties of each
individual species using standard kinetic theory expressions, and then determines the gas
mixture properties using mixture averaging rules. Note, in some cases CHEMKIN sub-
stitutes the mixture-averaged approach with a multicomponent approach to determine
the transport properties of the gas mixture. The interested reader is referred to the
CHEMKIN Theory Manual for further elaboration on the methodology and the ex-
pressions used in determining the ow properties of individual species and gas mixtures
[9].
In this study, priority is placed on determining the molecular transport parameters for
each species in the gas mixture, such that the CHEMKIN can determine the ow prop-
erties using its standard kinetic theory expressions. The molecular transport parameters
are inputted into CHEMKIN via a specied data format, as follows, in order:
1. An index indicating the geometrical conguration of the molecule. If the index is
0, then the molecule is a monoatomic. If the index is 1, then the molecule is linear.
If the index is 2, then the molecule is nonlinear.
2. The Lennard-Jones potential well depth, c,/
o
, in Kelvin.
3. The Lennard-Jones collision diameter, o, in angstroms.
4. The dipole moment, j, in Debyes (10
18
c:
32
c:p:
12
).
5. The polarizability, c, in cubic angstroms
6. The rotational relaxation collision number, 2
c|
, at 298 K.
The molecular transport parameters can be obtained from a variety of sources, and
CHEMKIN itself contains a transport database of over 200 species. However, more
species are often required when dealing with new fuels. The modeler can then turn to
previous modeling work to search for transport parameters. Alternatively, much data can
be found in standard reference texts such as Molecular Theory of Gases and Liquids
Chapter 3. Modeling Combustion Chemistry 26
[17] or other chemistry handbooks. If the data is still not found, then estimation and
analogy with related molecules can be used.
For new molecules, the Lennard-Jones collision diameter and potential well depth can
be estimated via dierent methods. Svehla describes how these parameters are obtained
by computing best ts to experimental data of a macroscopic transport property (e.g.,
viscosity) [18]. When experimental data is not available, as is the case for most gases,
these parameters are estimated using a variety of techniques. Svehla describes how
they can be calculated using the physical-chemical properties (e.g., boiling point, molar
volume, etc.), several empirical or combining rules, or other theoretical relations [18].
This study uses the correlations developed by Tee, Gotoh, and Stewart [19] and described
in Wang and Frenklach [20]. The correlations allow for the calculation of the Lennard-
Jones collision diameter and potential well depth using the critical pressure (P
c
) and
critical temperature (T
c
) of the gas. The critical temperature of a substance is the
temperature at and above which separate gas and liquid phases do not exist, and only
the supercritical state exists. The critical pressure is the vapor pressure at the critical
point (refer to Figure 3.1).
Figure 3.1: A typical phase diagram showing critical point
The equations for calculating the Lennard-Jones collision diameter and potential well
depth are given below. The accentric factor . in Equations 3.13 and 3.14 is evaluated
Chapter 3. Modeling Combustion Chemistry 27
using the Lee-Kesler vapor-pressure relations shown in Equation 3.15 [21]. The critical
temperature, critical pressure, and boiling point (T
o
) used in the calculations are obtained
from the NIST Chemistry WebBook [13]. If P
c
, T
c
, and (T
o
) are not readily available
for the species, they can be approximated from species with similar molecular structures.
o
(
1
c
1
c
)
13
= 2.3551 0.0874. (3.13)
o
c
/
o
1
c
= 0.7915 + 0.1693. (3.14)
. =
|:(1
c
) 5.927 + 6.096
T

T
1
+ 1.289|:
T

T
0.169
T

T
6
15.252 15.688
T

T
1
13.472|:
T

T
+ 0.436
T

T
6
(3.15)
The dipole moment (j) is a measure of the extent of polarity in covalent molecules.
It is dependent on the dierence electronegativity of the bonding atoms, and is precisely
dened as the product of the magnitude of the charge and the distance between the
charges. Nonpolar compounds, such as fully saturated hydrocarbons have zero dipole
moments while oxygenated compounds display higher dipole moments. Many experi-
mentally measured dipole moments are available in McClellans Tables of Experimental
Dipole Moments [22]. If experimental data is not available then the the molecular dipole
moment, which is a vector property, can be calculated for using vector addition of known
bond moments [23]. Such a method requires detailed information of the geometry of
molecular bonds and their electronegativities.
The polarizability (c) of a molecule quanties the tendency of a molecules charge
distribution (i.e., electron cloud) to be distorted from its normal shape by an external
electric eld (e.g., a nearby dipole or ion). Experimentally measured polarizability values
in cubic Angstroms can be obtained from the CRC Handbook of Chemistry and Physics
[24]. Bosque and Sales [25] have presented an empirical additive formula that allows the
estimation of polarizability from the molecular formula (i.e., # of C, H, and O atoms),
as shown in Equation 3.16.
c = 0.32 + 1.51 #C + 0.17 #H + 0.51 #C (3.16)
Chapter 3. Modeling Combustion Chemistry 28
3.4 Developing Chemical Kinetic Mechanisms
The process for developing and validating chemical kinetic mechanism was outlined by
Frenklach et al. [26] and summarized by Simmie [8]. Figure 3.2 is a owchart of the
mechanism development process. This owchart indicates that developing and validat-
ing a mechanism is a continuously evolving process wherein experiments and modeling
symbiotically achieve a satisfactory mechanism. The process can be summarized as fol-
lows:
1. Generate a list of elementary reactions.
2. Determine reaction rate constants for each reaction using literature sources or es-
timation, paying attention to temperature and pressure dependencies. Provide
thermochemical data to calculate equilibrium reverse rate constants.
3. Conduct controlled experiments that can be used to validate the reactions and rate
parameters given in the model.
4. Solve the reaction mechanism kinetics and transport equations using a computer
simulation of the experimental conguration. Conduct a sensitivity analysis to
determine the impact of specied rate constants on the nal result.
5. Compare the experimental data to the model predicted values. Optimize reac-
tion rate parameters that have the greatest impact on tting desired experimental
values.
The comprehensiveness of a mechanism is measured by its ability to describe com-
bustion phenomenon extensively. A mechanism is not considered comprehensive if it has
been tested against a single experiment because the role of each elementary reaction varies
with temperature, pressure, and composition. For example, reactions between hydrogen
atoms and fuel molecules are dominant in fuel-rich conditions, while reactions between
hydroxyl radicals and fuel molecules dominate in fuel-lean conditions. Many reactions
are important only at low temperatures, while others are dominant at high tempera-
tures. In an early treatise on chemical kinetic mechanisms for hydrocarbon combustion,
Westbrook [5] explains that a comprehensive mechanism must be validated against ex-
perimental data covering chemically reacting ows at various temperatures, pressures,
and reactant compositions.
Chapter 3. Modeling Combustion Chemistry 29
Determine
thermo-
chemical &
transport
properties
Conduct
controlled
validation
experiments
Simulate
experiments
using
computer
model
Compare
experiments
with computer
simulation
Formulate
elementary
reaction set
and determine
reaction rates
Figure 3.2: Flowchart for developing and validating chemical kinetic mechanisms
Autoignition characteristics are typically studied in shock tubes or rapid compression
machines, while reactions in a ameless premixed environment are studied in a jet stirred
reactor (i.e., perfectly stirred reactor). Laminar premixed and non-premixed ames are
often used to study combustion kinetics occurring at high temperatures. Each apparatus
can be operated at various temperature and pressure regimes to determine the eects of
these parameters on kinetic processes.
Since combustion of hydrocarbon fuels consists of sequential fragmentation of fuel
molecules into intermediates species, a comprehensive mechanism for any fuel must con-
tain detailed sub-mechanisms for the fuels intermediates. For example, since hydrogen
and carbon monoxide are products of hydrocarbon combustion, any hydrocarbon mecha-
nism must include reaction sub-mechanisms for hydrogen and carbon monoxide [7]. This
observation allows for the systematic and hierarchical development of kinetic mechanisms
by sequentially incorporating new species and reaction schemes in order of increasing
complexity [5].
Chapter 3. Modeling Combustion Chemistry 30
3.4.1 Mechanisms for Hydrocarbon Fuels
Chemical kinetic mechanisms for hydrocarbon fuels have been the focus of intense re-
search for several decades. Chemical kinetic mechanisms for biofuels, such as ethanol
and biodiesel, have only received attention recently. This section provides a background
on the combustion pathways of alkanes and alkenes because the same reaction types and
classes can be applied to biofuel combustion. In addition, detailed reaction rate stud-
ies for hydrocarbon fuels can be used to determine rate constants for similar reactions
for which no detailed reaction rate studies exist. Readers that are interested in a de-
tailed discussion of hydrocarbon combustion are directed towards recent review articles
by Battin-Leclerc [4] and Simmie [8].
A clear and simple explanation of the oxidation of fuels is given by Glassman [27].
Combustion reactions are driven by the formation of highly reactive radical, such as

O,

OH, and

H. During combustion, fuels are oxidized by a series of chain reactions which
can be categorized as one of following:
1. chain initiating,
2. chain propagating and chain branching,
3. chain terminating.
Chain initiating occurs when radical species are produced by dissociation of the re-
actants. The chain is propagated and branched as radicals react with stable compounds
to form additional radical species. Finally, the chain terminates when two radicals re-
combine to form stable species. The following subsections describe the major reaction
pathways for the combustion of alkanes and alkenes under low and high temperature
conditions. Comprehensive mechanisms would contain a number of minor reactions;
however, for the sake of simplicity, they have not been included here.
Combustion of Alkanes
Low Temperature Combustion of Alkanes
The combustion of hydrocarbons is dierent at low temperatures than high tempera-
tures. The general mechanism for the low temperature combustion of hydrocarbons was
developed by Semenov [28]. Benson [29] introduced the isomerization reaction of large
Chapter 3. Modeling Combustion Chemistry 31
hydrocarbons (Equation 3.20) to the Semenov mechanism. The following is a simplied
form of the Semenov mechanism including isomerization of large hydrocarbons:
1H +C
2
1 +HC
2
(3.17)
1 +C
2
c|/c:c + HC
2
(3.18)
1 +C
2
1C
2
(3.19)
1C
2
1CCH (3.20)
1CCH +C
2
1C +CH (3.21)
HC
2
+1H H
2
C
2
+1 (3.22)
H
2
C
2
+` CH +CH +` (3.23)
The chain is initiated by low temperature combustion of the hydrocarbon (1H) to
form an alkyl radical (1) and a hydroperoxy radical (HC
2
), as shown in Equation 3.17.
Next, the chain is propagated by one of two parallel reactions between alkyl radicals
and oxygen to form an alkene, HO
2
, and RO
2
(Equations 3.18 and 3.19). These reac-
tions compete with each other depending on the temperature. At temperatures above
500 K, Equation 3.18 predominates, wherein the oxygen abstracts a hydrogen from the
alkyl radical to form an alkene and a hydroperoxy radical. At temperatures below 500
K, Equation 3.19 is favored, wherein the oxygen adds to the alkyl radical to form an
alkylperoxy radical.
At low temperatures (below 500 K), propagation continues by the isomerization of
RO
2
to produce peroxide species (ROOH) (Equation 3.20). The radical pool then builds
up by degenerate branching of ROOH to form RO and OH radicals (refer to Equation
3.21). Further developments on this low temperature mechanism have been published by
Zhao et al. [30].
At intermediate temperatures (above 500 K), the HO
2
radical is more abundant,
so the reaction is propagated by hydrogen abstraction on the hydrocarbon by HO
2

to form hydrogen peroxide (H


2
O
2
) and an alkyl radical (refer to Equation 3.22). As
the temperature increases, hydrogen peroxide decomposes to form two hydroxyl radicals
(refer to Equation 3.23). The fuel-air mixture explodes once the radical pool builds up,
and then high temperature combustion predominates.
Chapter 3. Modeling Combustion Chemistry 32
Intermediate and High Temperature Combustion of Alkanes
The intermediate and high temperature combustion (e.g., above 900 K) of alkanes larger
than methane proceeds via either unimolecular decomposition or H-atom abstraction.
Unimolecular decomposition involves the breaking of the fuels C-C and/or C-H bonds,
and such reactions are typically favored at very high temperatures (e.g., above 1300-1400
K) and fuel rich conditions. The breaking of C-C bonds is favored over the breaking of
C-H bonds because of the bond dissociation energy is lower, and the reaction proceeds
as shown in Equation 3.24.
1H + (`) 1

+1

+(`) (3.24)
where
RH is an alkane molecule
R

and R

are alkyl radicals such as CH


3
, C
2
H
5
, etc.
(M) is a non-reacting collision partner
When highly reactive radicals, such as

O,

OH, and

H, are present they can abstract
H atoms from the fuel, as shown in Equation 3.25. These reactions are favored at
intermediate temperatures (e.g., 900-1300 K) and fuel lean conditions. Relative rate
coecients for H abstraction by radicals from tertiary, secondary, and primary CH bonds
are given in Table 3.1 [27]. Tertiary CH bonds are those on a carbon atom connected
to three other carbon atoms. Secondary CH bonds are on a carbon atom connected to
two other carbons. A primary CH bond is one on a carbon connected only to one other
carbon, such as the carbon at the end of a hydrocarbon chain. The table indicates that
tertiary CH bonds are the weakest and primary CH bonds are the strongest. If the fuel is
an alkene then radicals will abstract H from carbon atoms that are singly bonded because
CH bonds on doubly bonded carbon atoms are very strong.
1H + A 1 +AH (3.25)
where X is any radical specie, usually O, OH, H, and CH
3

The alkane radical then decays to form an alkene and a radical specie, as shown
in Equation 3.26. For example, the isopropyl radical, obtained from H abstraction of
the secondary CH bond in propane, will decay to propene and a H atom, as shown in
Chapter 3. Modeling Combustion Chemistry 33
Table 3.1: Relative magnitudes of rate constants for H abstraction from dierent CH
bonds [27]
Tertiary Secondary Primary
H 13 4 1
O 10 5 1
HO
2
10 3 1
OH 4 3 1
Figure 3.3. The process by which the alkyl radical decomposes is called -scission. In -
scission, the bond once removed from the radical site will break to form an alkene without
a hydrogen shift. Furthermore, C-C bonds are more likely break than C-H bonds since
C-C bonds are weaker.
1 c|/c:c + 1

(3.26)
where
R

is a hydrocarbon radical or H atom


Figure 3.3: Decay of isopropyl radical
Combustion of Alkenes
Alkane combustion ends with the formation of alkenes and a pool of radical species, so
the combustion of these alkene compounds will now be discussed, taking ethene as an
example. First, the C=C double bond is attacked primarily by the biradical O , which
forms an intermediate species that subsequently decays, as shown in Figure 3.4. Thus,
the two primary addition reactions are shown in Equations 3.27 and 3.28. Some minor
Chapter 3. Modeling Combustion Chemistry 34
reactions involving H abstraction by OH and H radicals also play a role, as shown in
Equations 3.29 and 3.30, respectively.
C
2
H
4
+ C CH
3
+HCC (3.27)
C
2
H
4
+ C CH
2
+CH
2
C (3.28)
C
2
H
4
+CH C
2
H
3
+H
2
C (3.29)
C
2
H
4
+ H C
2
H
3
+H
2
(3.30)
Figure 3.4: Addition of O radical to ethene from [27]
Then the vinyl radical (C
2
H
3
) decays to acetylene, as shown in Equation 3.31. The
acetylene is consumed by a reaction with the biradical O to form a methylene radical and
carbon monoxide, as shown in Equation 3.32. The fate of CH
3
, CH
2
O (formaldehyde),
CH
2
, and CO are described in the mechanism for methane [27].
C
2
H
3
+` C
2
H
2
+ H +` (3.31)
C
2
H
2
+C CH
2
+CC (3.32)
Mechanism of Soot Formation
The term soot refers to tiny amorphous carbon particles produced from the combustion
of hydrocarbon fuels. Soot emissions have become an environmental concern due to the
negative impacts of particulate matter on the human respiratory system. In addition, soot
particles in the atmosphere can contribute to global warming by altering the radiative
balance of the atmosphere [31].
Chapter 3. Modeling Combustion Chemistry 35
The presence of soot particles in hydrocarbon ames is identiable by their char-
acteristic yellow-orange appearance. This color is generated by photons emitted from
the solid carbon particulates. On the other hand, ames that do not contain soot are
blue in color. The formation of soot particles in a combustion environment is a highly
complex mechanism and depends on a number of factors, e.g. fuel composition, temper-
ature, fuel-oxygen ratio, ame conguration, etc.. However, the formation of soot has
been shown to be highly controlled by chemistry related phenomenon [32]. Glassman
[27] has summarized the work of other researchers to describe the dominant route of soot
formation.
Initially, the fuel breaks down to acetylene, as mentioned in the previous discussion
on hydrocarbon oxidation. In the high-temperature post-ame regime, soot formation is
initiated by the growth of small straight-chain alkenes (acetylene) to small aromatic com-
pounds (e.g. benzene). The aromatic hydrocarbons then react sequentially with smaller
hydrocarbons (acetylene, in particular) to form larger polyaromatic hydrocarbon (PAH)
species. Gaseous PAH molecules continue to nucleate until the smallest identiable soot
particles appear, with diameters of a few nanometers.
Figure 3.6 [27] shows the mechanism for soot formation in more detail. Initially,
the acetylene (C
2
H
2
) undergoes H addition to form the vinyl radical. The vinyl radical
then reacts with another acetylene molecule to form the 1,3-butadienyl radical. The
1,3-butadienyl radical can also be readily formed from C
4
hydrocarbons via hydrogen
abstraction then -scission.
In diusion ames, the alternate route A is then followed, wherein the 1,3-butadienyl
radical reacts again with acetylene to form the cyclic phenyl (C
6
H
5
) radical, following ring
closure. The phenyl radical is essentially a benzene molecule missing one hydrogen atom.
The phenyl radical is also produced by alternate route C, wherein methyl acetylene (i.e.,
propyne C
3
H
4
) pyrolyzes rapidly to form the aromatic. The phenyl radical can proceed
to grow into a larger aromatic via the two-step hydrogen abstraction carbon addition
(HACA) mechanism. In the HACA mechanism, the aromatic molecule is converted to
a radical by hydrogen abstraction, and then grows in size by the (carbon) addition of
an acetylene molecule. The HACA mechanism continues until larger PAH molecules
appear. As the concentration of gaseous PAH species increases, nucleation occurs and
soot particles begin to appear.
This thesis study does not directly measure the soot levels generated in biofuel diu-
sion ames. However, the concentrations of many soot precursors are measured, speci-
Chapter 3. Modeling Combustion Chemistry 36
cally, acetylene, C
4
hydrocarbons, 1,3-butadiene, and benzene. Analyzing the concentra-
tions of these species in the diusion ame can oer an insight into the sooting potential
of various biofuel fuels.
3.4.2 Determining Rate Coecients
The ideal method for determining reaction rate coecients for each reaction in the mech-
anism is to conduct fundamental experiments over the range of temperatures encountered
during combustion (e.g., typically 300-2500 K). However, such experiments are dicult
and have not been performed for many of the fuel and intermediate species encountered
in biofuel combustion. Computational methods based on high-level ab initio quantum
chemistry and density function theory (DFT) are also available for determining rate coef-
cients. However, ab initio methods are computationally expensive while DFT methods
are considered inaccurate [33, 34]. A simpler heuristic method of determining reaction
rate constants uses analogies to similar well studied reactions [35].
Determining rate constants using analogies requires model compounds of similar
structure and chemical bonding characteristics. Sucient experimental and theoreti-
cal information must be available for the model compounds to make the analogy valid.
Figure 3.5 provides the chemical structure and bond dissociation energy (BDE) [24] of
some common alkanes and alcohols, and this information is used to display how analogies
can be used to determine reaction rate constants for n-butanol combustion.
First, it is observed that the chemical structure of n-butanols functional group is
similar to that of ethanol because both are alcohols. The BDE of the O-H bond in the
hydroxyl group of ethanol and n-butanol is roughly the same, given the reported error
margins. Therefore, reactions involving H-atom abstraction from the hydroxyl group
in n-butanol would have the same reaction rate constant as the analogous reaction in
ethanol. The BDE between the c C and the OH group is also the same in n-butanol
and ethanol; therefore, scission of C-OH bond in butanol would have the same reaction
rate as the analogous reaction in ethanol. The BDE of the c C-H bond is not reported
for butanol. However, the BDE of ethanols c C-H bond is observed to be nearly the
same as the BDEs of the central C-H bonds in propane and n-butane, and therefore
it is likely that butanols analogous bond has a similar BDE. Reactions involving H-
atom abstraction from n-butanols c C are likely to have reactions rates identical to the
analogous reaction in ethanol. The BDE for breaking the c and carbons in n-butanol is
Chapter 3. Modeling Combustion Chemistry 37
close to the BDE of the analogous bond in n-butane, ethanol, and propane, and therefore,
it is possible to use analogies to determine the the reaction rate constants. The BDE for
the remaining , , and o carbons in n-butanol are unknown, but they are likely similar
to those of n-butane because the destabilizing eects of the hydroxyl group, which are
only strong on the adjacent atoms, are minimal.
C H
3
C
H
2
C
H
2
C
H
2
OH
C H
3
C
H
2
OH
C H
3
C
H
2
C
H
2
CH
3
C H
3
C
H
2
CH
3
432.3 kJ/mol 5 kJ/mol
389.9 4.2 kJ/mol
441 5.9 kJ/mol 421.7 8 kJ/mol
401.2 4.2 kJ/mol
391.2 2.9 kJ/mol
357.3 3 kJ/mol
364.8 4.2 kJ/mol
n-butanol
ethanol
411 .1 2 .2 kJ/mol
421 .3 kJ/mol
363.2 2.5 kJ/mol
372 2.9 kJ/mol
411 .1 2 .2 kJ/mol
n-butane
422 .2 2 .1 kJ/mol
410 .5 2 .9 kJ/mol
379.3 2.1 kJ/mol
379.3 2.1 kJ/mol
422 .2 2 .1 kJ/mol
propane
Figure 3.5: Chemical structures and bond dissociation energies for alcohols and alkanes
[24]
Chapter 3. Modeling Combustion Chemistry 38
F
i
g
u
r
e
3
.
6
:
G
e
n
e
r
a
l
m
e
c
h
a
n
i
s
m
f
o
r
s
o
o
t
f
o
r
m
a
t
i
o
n
f
r
o
m
G
l
a
s
s
m
a
n
[
2
7
]
Literature Cited
[1] J. Heywood, Internal Combustion Engine Fundamentals. McGraw-Hill Book Com-
pany, New York., 1988.
[2] A. A. Amsden, J. D. Ramshaw, P. J. ORourke, and J. K. Dukowicz, KIVA a com-
puter program for two- and three-dimensional uid ows with chemical reactions and
fuel sprays, Los Alamos National Laboratory, Tech. Rep. LA-10245-MS, February
1985.
[3] V. Basevich, Chemical-kinetics in the combustion processes - a detailed kinetics
mechanism and its implementation, Progress in Energy and Combustion Science,
vol. 13, no. 3, pp. 199248, 1987.
[4] F. Battin-Leclerc, Detailed chemical kinetic models for the low-temperature com-
bustion of hydrocarbons with application to gasoline and diesel fuel surrogates,
Progress in Energy and Combustion Science, vol. 34, no. 4, pp. 440498, August
2008.
[5] C. K. Westbrook and F. L. Dryer, Chemical kinetic modeling of hydrocarbon com-
bustion. Progress in Energy and Combustion Science, vol. 10, no. 1, pp. 1 57,
1984.
[6] R. J. Kee, M. E. Coltrin, and P. Glarborg, Chemically Reacting Flow: Theory and
Practice. Wiley, 2005.
[7] T. Lu and C. K. Law, Toward accommodating realistic fuel chemistry in large-
scale computations, Progress in Energy and Combustion Science, vol. 35, no. 2, pp.
192215, April 2009.
39
Literature Cited 40
[8] J. M. Simmie, Detailed chemical kinetic models for the combustion of hydrocarbon
fuels, Progress in Energy and Combustion Science, vol. 29, no. 6, pp. 599 634,
2003.
[9] R. J. Kee, F. M. Rupley, J. A. Miller, M. E. Coltrin, J. F. Grcar, E. Meeks, H. K.
Moat, A. E. Lutz, G. Dixon-Lewis, M. D. Smooke, J. Warnatz, G. H. Evans,
R. S. Larson, R. E. Mitchell, L. R. Petzold, W. C. Reynolds, M. Caracotsios, W. E.
Stewart, P. Glarborg, C. Wang, C. L. McLellan, O. Adigun, W. G. Houf, C. P. Chou,
S. F. Miller, P. Ho, P. D. Young, D. J. Young, D. W. Hodgson, M. V. Petrova, and
K. V. Puduppakkam, CHEMKIN Theory Manual, 4th ed. San Diego, California:
Reaction Design, June 2006.
[10] H. McDonald, Combustion modeling in 2 and 3 dimensions - some numerical con-
siderations, Progress in Energy and Combustion Science, vol. 5, no. 2, pp. 97122,
1979.
[11] R. J. Kee, F. M. Rupley, J. A. Miller, M. E. Coltrin, J. F. Grcar, E. Meeks, H. K.
Moat, A. E. Lutz, G. Dixon-Lewis, M. D. Smooke, J. Warnatz, G. H. Evans,
R. S. Larson, R. E. Mitchell, L. R. Petzold, W. C. Reynolds, M. Caracotsios, W. E.
Stewart, P. Glarborg, C. Wang, C. L. McLellan, O. Adigun, W. G. Houf, C. P. Chou,
S. F. Miller, P. Ho, P. D. Young, D. J. Young, D. W. Hodgson, M. V. Petrova, and
K. V. Puduppakkam. (2006) CHEMKIN release 4.1. San Diego, California.
[12] Anonymous. (2009, January) NIST chemical kinetics database, standard reference
database 17, version 7.0 (web version), release 1.4.3.
[13] P. Linstrom and W. Mallard. (2005) NIST chemistry webbook, NIST standard
reference database number 69. National Institute of Standards and Technology,.
Gaithersburg, Maryland, 20899. [Online]. Available: http://webbook.nist.gov
[14] S. Benson, Thermochemical Kinetics, 2nd edition. Wiley, New York, 1976.
[15] C. Muller, V. Michel, G. Scacchi, and G. Come, THERGAS - a computer-program
for the evaluation of thermochemical data of molecules and free-radicals in the gas-
phase, Journal de Chimie Physique et de Physico-Chimie Biologique, vol. 92, no. 5,
pp. 11541178, May 1995.
Literature Cited 41
[16] E. R. Ritter and J. W. Bozzelli, THERM: Thermodynamic property estimation
for gas phase radicals and molecule, International Journal of Chemical Kinetics,
vol. 23, pp. 767778, 1991.
[17] J. O. Hirschfelder, C. F. Curtiss, and R. B. Bird, Molecular Theory of Gases and
Liquids. Wiley, New York, 1964.
[18] R. Svehla, Estimated viscositites and thermal conductivities of gases at high tem-
peratures, National Aeronautics and Space Administration, Tech. Rep., 1962.
[19] L. Tee, S. Gotoh, and W. Stewart, Molecular parameters for normal uids - lennard-
jones 12-6 potential, Industrial & Engineering Chemistry Fundamentals, vol. 5,
no. 3, p. 356, 1966.
[20] H. Wang and M. Frenklach, Transport properties of polycyclic aromatic hydrocar-
bons for ame modeling, Combustion and Flame, vol. 96, no. 1-2, pp. 163170,
January 1994.
[21] B. Lee and M. Kesler, A generalised thermodynamic correlation based on the three-
parameter corresponding states, AIChE Journal, vol. 21, pp. 510527, 1975.
[22] A. McClellan, Tables of Experimental Dipole Moments. San Francisco: Freeman,
1963.
[23] S. Bohm and O. Exner, Prediction of molecular dipole moments from bond mo-
ments: testing of the method by DFT calculations on isolated molecules, Physical
Chemistry Chemical Physics, vol. 6, no. 3, pp. 510514, 2004.
[24] D. R. Lide, Ed., CRC Handbook of Chemistry and Physics, 87th Edition. Boca Ra-
ton, FL: Taylor and Francis, 2007. [Online]. Available: http:/www.hbcpnetbase.com
[25] R. Bosque and J. Sales, Polarizabilities of solvents from the chemical composition,
Journal of Chemical Information and Computer Sciences, vol. 42, no. 5, pp. 1154
1163, September-October 2002.
[26] M. Frenklach, H. Wang, and M. Rabinowitz, Optimization and analysis of large
chemical kinetic mechanisms using the solution mapping method: combustion of
methane. Progress in Energy and Combustion Science, vol. 18, pp. 4773, 1992.
Literature Cited 42
[27] I. Glassman, Combustion, 3rd ed. San Diego, CA: Academic Press, 1996.
[28] N. N. Semenov, Some Problems in Chemical Kinetics and Reactivity. Princeton,
New Jersey: Princeton University Press, 1958.
[29] S. W. Benson, Kinetics and thermochemistry of chemical oxidation with application
to combustion and ames. Progress in Energy and Combustion Science, vol. 7, no. 2,
pp. 125 134, 1981.
[30] F. Zhao, T. W. Asmus, D. N. Assani, J. E. Dec, J. A. Eng, and P. M. Najt, Ho-
mogeneous Charge Compression Ignition (HCCI) Engines: Key Research and De-
velopment Issues. Warrendale, PA, USA: Society of Automotive Engineers Inc.,
2003.
[31] T. C. Bond and K. Sun, Can reducing black carbon emissions counteract global
warming? Environmental Science and Technology, vol. 39, no. 16, pp. 5921 5926,
2005.
[32] I. M. Kennedy, Models of soot formation and oxidation, Progress in Energy and
Combustion Science, vol. 23, no. 2, pp. 95 132, 1997.
[33] J. Senosiain, J. Han, C. Musgrave, and D. Golden, Use of quantum methods for
a consistent approach to combustion modelling: Hydrocarbon bond dissociation
energies, Faraday Discussions, vol. 119, pp. 173189, 2001.
[34] C. J. Hayes and D. R. Burgess, Jr., Exploring the oxidative decompositions of
methyl esters: Methyl butanoate and methyl pentanoate as model compounds for
biodiesel, Proceedings of the Combustion Institute, vol. 32, no. Part 1, pp. 263270,
2009.
[35] N. Marinov, A detailed chemical kinetic model for high temperature ethanol ox-
idation, International Journal of Chemical Kinetics, vol. 31, no. 3, pp. 183220,
March 1999.
Chapter 4
Experimental Apparatus and
Analytical Methodology
This section provides a detailed description of the experimental setup used to study
opposed-ow diusion ames of biofuels. The setup has been previously described in the
MASc thesis entitled Using an opposed-ow diusion ame to study the oxidation of C
4
fatty acid methyl esters by S.M. Sarathy [1]. The setup was modied in order to study
the proposed biofuels, so this chapter presents a description of the modied setup. This
chapter does not include descriptions of other experimental setups (i.e. the jet stirred
reactor) used to validate chemical kinetic mechanisms for biofuels, since these setups are
not the authors contribution.
The setup is designed to generate an opposed-ow diusion ame from a liquid or
gaseous fuel. The concentrations of stable species and the temperature prole in the
ame are then obtained. The setup consists of the fuel delivery system, the opposed
ow diusion ame burner, the sample collection apparatus, and a variety of analytical
instruments. The fuel delivery system pumps the liquid fuel into a vaporizing chamber
that produces a mixture of vaporised fuel and nitrogen. The gaseous fuel stream and
the oxidizer stream then ow into an opposed ow burner to produce a planar ame.
Samples are extracted from the ame region using a fused silica micro-probe connected
to a vacuum pump. Analysis of hydrocarbon compounds is performed using a gas chro-
matograph ame ionization detector (GC/FID). Carbon dioxide and carbon monoxide
are quantied by an non-dispersive infrared (NDIR) analyzer. The ame temperature
prole is measured using an R-type thermocouple.
43
Chapter 4. Experimental Apparatus and Analytical Methodology 44
4.1 Opposed-ow Diusion Burner Setup
The opposed-ow diusion ame oers experimental and modeling advantages over co-
ow ames, although their stability is more sensitive to ow conditions [2]. The opposed-
ow conguration results in a planar ame which simplies ame analysis to a one-
dimensional system. The species concentration and temperature are a function of axial
distance only. Furthermore, the modeling of uid mixing is simplied in laminar ames
because fuel and oxidizer mixing is limited to diusional processes (i.e., turbulent mixing
patterns are not considered).
These experiments utilize two identical burners
1
with circular burner ports. The two
burner ports are placed opposite each other in the same vertical plane. A fuel mixture
is fed through the bottom port while an oxidizer mixture is fed through the top. The
two opposing streams ow into each other to create a stagnation plane between the two
ports. The vertical location of the stagnation plane depends on the momentum of the two
streams. The stagnation plane prevents any non-diusional mixing between the fuel and
oxidizer. As the two streams molecularly diuse into each other, a at ame is ignited.
The ame is physically located wherever the stoichiometric mixture fraction of fuel and
oxygen exists. Therefore, the exact location depends on the mole fractions of fuel and
oxygen in their respective streams, as well as each streams diusivity.
A diagram of the burner port is shown in Figure 4.1. Each burner port consists of a
stainless steel housing enclosing a porous sintered bronze matrix. The porous material
is divided into inner and outer coaxial cylinders of diameter 25.4 mm and 38.1 mm,
respectively. The inner cylinder directs the fuel or oxidizer streams towards each other,
and the outer annulus can be used to create a nitrogen shroud around the ame for
minimizing external ow disturbances. The annulus shrouding feature was not exploited
in this study. Plug ow boundary conditions were assumed for the modeling of this setup.
The temperature of the gases owing through the ports was controlled by circulating a
heat transfer uid through porous plugs packed between two co-axial cylinders. This
ensures that the gases have a uniform laminar ow and at velocity prole at the ports
surface. The heat transfer uid was maintained at set temperature using either a water
heating recirculator
2
or an electric air heater
3
. Water was the heat transfer uid for
1
Purchased from Holthuis & Associates McKenna Flat Flame Burners
2
PolyScience Heating Recirculator Model 210
3
Omega process heater
Chapter 4. Experimental Apparatus and Analytical Methodology 45
experiments requiring burner port temperatures below 100

C, while air was the heat
transfer uid in experiments requiring burner port temperatures above 100

C (refer to
Table 4.2). Additional heat input is also provided by wrapping the bottom port with
heating tapes
4
.
Figure 4.1: Diagram of burner port
A photograph of the actual burner setup is shown Figure 4.2. The two burners
are coaxially mounted 20 mm apart facing each other. A custom-built aluminum holder
provides support to the burners, and a clear quartz shroud 22 cm in diameter protects the
ame from airow disturbances from the surrounding environment. The entire assembly
is mounted atop a translation stage
5
which moves along the vertical axis with the rotation
of a micrometer knob. This made it possible to obtain a vertical prole of the ame
temperature and emissions between the two burner ports. Combustion products from
the burner ow upwards into a ventilated hood which is ducted to the laboratory fume
hood.
4.2 Fuel Preparation and Vaporization
The fuel stream fed to the burner is a mixture of the fuel and nitrogen gas. Properties
of each fuel are shown in the following Table 4.1, which indicates that three of the fuels
4
Omega FGS Standard Insulated High Temperature Heating Tapes
5
Newport M-MVN120 Precision Ball Bearing Vertical Linear Stage, 20mm Travel
Chapter 4. Experimental Apparatus and Analytical Methodology 46
Figure 4.2: Photograph of burner setup
are liquids at room temperature. Gaseous fuels are stored in compressed gas cylinders
and mixed in line with nitrogen before being delivered to the burner. Liquid fuels are
pumped from a bottle to an ultrasonic atomizer. The atomizer probe sprays the fuel into
a heated stainless steel mixing chamber where it mixes with nitrogen gas. The mixture
of fuel and nitrogen is then fed to the bottom burner port.
Table 4.1: Physical Chemical Properties of the Fuels Used
Mol. Wt. (g/mole) Phase at STP Normal Boiling Pt. (

C)
n-butanol 74.1 liquid 117
n-butane 58.1 gas 0
methyl decanoate 186.3 liquid 224
Liquid fuel is pumped from its stock bottle by a peristaltic pump head driven by
a motor
6
with a polytetrauoroethylene (PTFE) pump head. The pumping system is
calibrated for each fuel at the beginning of each experiment using a graduated cylinder
to measure the volume of liquid pumped at a various oump head drive rates. The pump
drive rate is then set to correspond with the desired volumetric ow rate. The selected
ow rates for each fuel are provided in the next section.
A new vaporization system designed by a postdoctoral fellow, Dr. John Z Wen, to
deal with the high boiling point liquid fuels is used in this study. The design consists of
a commercially available ultrasonic atomizer which is threaded to a heated stainless steel
6
Cole-Parmer Masterex L/S Microprocessor Pump Drive
Chapter 4. Experimental Apparatus and Analytical Methodology 47
chamber. The pumped fuel is delivered to a 30 kHz ultrasonic atomizer
7
unit that breaks
the fuel into micro-droplets. The units ultrasonic power supply converts 60 Hz energy to
a high frequency electrical energy at 30 kHz. Then, a piezoelectric transducer converts
the electrical energy to mechanical vibrations. The vibrations are intensied in a probe
and focused at its tip. The liquid fuel is dispensed in the probe, where it spreads out as
a thin lm on the tip. The oscillations at the tip atomizes the liquid into micro-droplets
to form a gentle, low viscosity mist. The body of the atomizer was kept cool at 30

C by
blowing compressed air over it.
A schematic of the fuel vaporization/mixing chamber is shown in Figure 4.3. The
atomized fuel is sprayed into the mixing chamber, where it mixes with nitrogen gas
8
.
The nitrogen was not preheated in this study because high temperatures near the top
of the mixing chamber can overheat the atomizer. The temperature in the lower half
of the chamber was maintained using heating tapes
9
and a thermal blanket
10
. A high
temperature in only the lower portion of the mixing chamber served to vaporize the fuel
droplets while minimizing the heat transfer to the atomizer. The temperature inside the
chamber is measured using a stainless steel sheathed K-type thermocouple inserted at the
top of the chamber. Specic details on the temperature set point are provided in Table
4.2 in the next section. The mixing chamber is a custom-made stainless steel column
11
30
cm in length and 10 cm in outer diameter (OD), with a 15 cm OD, 3,8

thick, stainless
steel ange bolted at the top to accommodate insertion of the atomizer, nitrogen, and
thermocouple. The column is packed with a 20 cm long bed of glass marbles. The
bed increases the path length traversed by gaseous fuel and nitrogen molecules; thereby,
increasing mixing of the two streams. The gaseous mixture of fuel and nitrogen then
ows to the bottom port of the burner apparatus via a 1,4

stainless steel transfer line


heated to 250

C
12
.
It was found that the atomizer did not operate properly when threaded onto a 3,8

thick stainless steel ange plate. The thick plate dampens the oscillations transmitted
to the atomizer tip and prevents the tip from vibrating. By working with the atomizer
manufacturer, it was determined that a portion of the anged plate should be machined
7
Sonaer custom built 30 kHz model
8
Linde Grade 4.8
9
Omega SWH Ultra-High Temperature Heating Tapes
10
Unifrax InsulfraxS blanket
11
Designed by Dr. John Z. Wen
12
Unique Heated Products Instrument Grade Heating Sample Line
Chapter 4. Experimental Apparatus and Analytical Methodology 48
to 1,32

thick, in order to minimize the dampening eects. The thin portion of the
machined ange plate allowed the atomizer to function properly while still meeting the
other design requirements of the ange. Refer to Figure 4.4 for a schematic of the
machined ange plate.
Chapter 4. Experimental Apparatus and Analytical Methodology 49
Figure 4.3: Schematic of the mixing chamber. Design by Dr. John Z. Wen
Figure 4.4: Schematic of the machined ange plate.
Chapter 4. Experimental Apparatus and Analytical Methodology 50
4.3 Supply of Fuel and Oxidizer Streams
The ow rates of fuel and oxidizer streams through the burner ports were key parameters
in these experiments. It is important that the momentums of the two streams be nearly
equal, so that a stagnation plane is created where the streams meet. The molar concen-
trations of fuel and oxidizer in each stream needs to be sucient enough to light a ame;
however, high sooting ames are not desired. The ow rates of nitrogen, oxygen and air
are controlled using mass ow controllers
13
, while the ow rate of liquid fuel is controlled
by a peristaltic pump. The mass ow controllers are regularly recalibrated using a pos-
itive displacement gas ow meter
14
. The inlet oxidizer and fuel stream concentrations
are selected based on the following criteria:
a low Reynolds Number to create a laminar ame
a low sooting ame to prevent clogging of sampling probe
avoid a very hot ame that will damage the probe
a balanced momentum of the two streams to form a stagnation plane at their
intersection
avoid excessive unburned fuel
at the ame plane, an N
2
/O
2
ratio near that of air to make the study relevant to
actual ames
The molar composition and measured temperature of the fuel and oxidizer streams
entering the burner for each experiment is indicated in Table 4.2. The temperature
of the oxidizer stream exiting the top burner was near 140

C. The cause of this high
temperature was heat convected towards the top burner port from rising combustion
product gases.
It is important to ensure that the liquid fuel is suciently vaporized in the mixing
chamber and that it does not condense before exiting the burner, which has a maximum
operating temperature of 130

C. To prevent condensation, the experiments conducted in
this study were performed at low partial pressures of fuel (i.e., 0.02-0.06 atm) wherein the
13
Teledyne-Hastings Mass Flow Controller HFC202
14
BIOS Dener 220
Chapter 4. Experimental Apparatus and Analytical Methodology 51
vaporisation temperature is depressed. The vapour pressure of dierent FAME species
[3] was plotted to determine which fuels could be used under the temperature constraints
of the burner. Figure 4.5 indicates that FAME with 10 carbons or less can be suciently
vaporised at 130

C.
Table 4.2: Experimental conditions
Mixing Chamber Fuel Stream Fuel Port Oxidizer Stream
T (

C) Composition Exit T (

C) Composition
n-butanol 110 5.9% Fuel 80 42.1% O
2
94.1% N
2
57.9% N
2
n-butane N/A 5.9% Fuel 30 42.1% O
2
94.1% N
2
57.9% N
2
methyl decanoate 150 1.8% Fuel 110 42.1% O
2
98.2% N
2
57.9% N
2
1.0E-05
1.0E-04
1.0E-03
1.0E-02
1.0E-01
1.0E+00
1.0E+01
1.0E+02
0 50 100 150 200 250 300
Temperature (C)
V
a
p
o
u
r

P
r
e
s
s
u
r
e


(
a
t
m
)
Methyl Hexanoate
Methyl Octanoate
Methyl Decanoate
Methyl Dodecanoate
Methyl Oleate
Figure 4.5: Vapor pressure curves for fatty acid methyl esters
Chapter 4. Experimental Apparatus and Analytical Methodology 52
4.4 Reynolds Number and Strain Rate Calculations
The Reynolds Number (1
c
) is used to determined whether the ow exiting the burner
port is laminar (1
c
< 2300), transient (2300 < 1
c
< 4000), or turbulent (1
c
4000).
The goal of these experiments is to generate a laminar diusion ame, so the uid velocity
must be set accordingly. Equation 4.1 is used to calculate the Reynolds Number for ow
in the burner port assuming a plug ow behaviour.
1c =
j n d
j
(4.1)
where 1c is the Reynolds Number
j is the gas density in
Ij
n
3
n is the gas velocity in
n
-
d is the burner port diameter in :
j is the dynamic viscosity in
.-
n
2
The strain rate is dened as the normal gradient of the normal component of the ow
velocity [4]. The strain is calculated using Equation 4.2.
c
1
=
2\
1

1
(
1 +
\
2

j
2
\
1

j
1
)
(4.2)
where c
1
is the strain rate on the fuel side in :
1
1 is the distance between the two burner ports in c:
\
1
is the absolute value of the fuel stream velocity at the fuel boundary
in
cn
-
\
2
is the absolute value of the oxidizer stream velocity at the oxidizer
boundary in
cn
-
j
1
is the fuel stream density in
j
cn
3
j
2
is the air stream density in
j
cn
3
Chapter 4. Experimental Apparatus and Analytical Methodology 53
4.5 Gas Sampling System
The previous sections discussed the procedure for producing an opposed ow diusion
ame. Once the ame was generated, a sampling system was used to obtain qualitative
and quantitative information about the ames characteristics. Specically, the species
concentrations at various points between the two burner ports was measured to obtain
characteristic proles. The following sections discuss the gas sampling system, as well as
the sampling procedure.
4.5.1 Sampling Apparatus
Microprobes, due to their small perturbation of ow elds, are commonly used in ame
studies to acquire the concentration of stable species [5, 6, 7, 8]. The gas sampling system
in these experiments consists of a quartz microprobe connected to a dual-stage pump
15
via a heated 1,4

stainless steel line and a vacuum pressure gauge. The microprobe is


mounted on a sliding stage, allowing it to move into and out of the ame region easily.
The rst stage (vacuum) of the pump creates a suction in the sampling line to withdraw
gas samples from the ame. An analytical instrument is connected downstream of the
second stage (compressor) of the pump to study the gases owing through the line. The
compressor head on the pump pushes samples into the analytical instrument via a 1,4

stainless steel transfer line heated to 250



C
16
. The heated transfer lines are required to
prevent the condensation of high molecular weight species out of the sample gas.
Previous studies have identied precautions to take when using the microprobe sam-
pling technique. The primary objective is to eliminate chemical reactions within the
probe and sampling lines. A combination of rapidly reducing temperature and pressure
in the probe helps meet this objective.
Kassem and coworkers [9] studied the eect of microprobe cooling on fuel-rich, laminar
at ames of chlorinated hydrocarbons. Their results indicate that cooled probes, as
opposed to uncooled probes, provided more accurate proles of species concentration.
Schoenung and Hanson [5] showed that carbon monoxide (CO) measurements in the
post-ame region of a premixed methane/air ame were aected by the pressure within
the probe and sampling lines. Their results indicate that the CO concentration increases
as the pressure in the probe decreases, with the concentration reaching the actual value
15
KNF oil-free dual-staged pump Model UN035.3 ST11 with heated heads
16
Unique Heated Products Instrument Grade Heating Sample Line
Chapter 4. Experimental Apparatus and Analytical Methodology 54
around 50 mm Hg. This nding suggests that CO is converted to CO
2
in the probe region
unless the pressure is 7 kPA or below. Therefore, low temperatures and pressures in the
probe are required to quench the reactions.
Fristrom and coworkers [10] argue that it is not rapid temperature drop that is re-
quired for successful ame sampling. Instead, a combination of rapid pressure drop and
the destruction of radicals on the probe walls is responsible for quenching reactions at
the probe tip. They explain that 2
ao
order molecule-radical reaction rates vary with
the square of the gas density, which varies linearly with pressure; therefore, the reac-
tion rate decreases with a decrease in pressure. Furthermore, they argue that if a rapid
temperature drop is induced while keeping pressure constant, then reactions with acti-
vation energies of 20 kJ/mol or lower will increase in rate. These reaction rates vary
quadratically with density, which varies inversely with temperature at constant pressure;
therefore, decreasing the temperature alone would not quench all ame reactions.
The error introduced during microprobe sampling of laminar ame species proles is
a concern. The insertion of a microprobe into a ame can disturb the uid ow, provide a
surface for quenching reactions, oer poor spatial resolution, and absorb thermal energy.
These eects are certainly not reproducible in modeling simulations, so some researchers
make corrections to the experimental data to compensate for errors that microprobes
introduce (e.g., shifting species proles, increasing measured concentrations, etc.). A
recent study by Struckmeier et al. [11] compared the quality of species proles obtained
using intrusive microprobes and nonintrusive optical techniques. The results indicate
that data acquired using both techniques agreed well, and therefore suggests that probe
sampling techniques are an indispensable technique for reliable experimental data. In
addition, the authors state that only minor corrections (e.g., shifting proles by 1 mm
and measured concentrations by 30%) should be made to compensate for errors in the
experimental data.
One set of experiments in this study used a quartz microprobe fabricated in the
Department of Chemistry Glass Blowing Shop. The inner diameter (ID) of the probe
tip could not be precisely controlled; therefore, a number of probes were made and
measurements were taken to select the best one. To avoid errors associated with probe
fabrication and the subsequent measurement procedures, a new sampling probe apparatus
was also designed as described below. A previous study by Syed [12] determined the
appropriate probe tip size. The study measured CO
2
concentrations in a propane-air
ame using probe tips with various IDs. Syed suggests that a probe tip with an ID
Chapter 4. Experimental Apparatus and Analytical Methodology 55
of approximately 150-250jm is ideal. Probes with larger IDs do not successfully quench
reactions at the probe tip. Probes with smaller IDs restrict gas ow through the sampling
line, such that it is not possible to obtain gas samples from the ame.
In addition to having a small ID, the probe tip must also be long enough to allow
for a large pressure drop. As mentioned previously, the microprobe is connected to a
doubled-headed pump via 1,4

tubing. If the suction pump head creates an absolute


pressure of 4-6 kPa in the sampling line
17
, then this is sucient to quench reactions. At
such low sampling pressures, the probe is not cooled since previous studies have shown
that a rapid pressure drop alone provides accurate sampling [10, 5]. This study found
that a probe tip ID of 250jm and a length of 3.5-4.0 cm achieves the aforementioned
pressure range in the sampling line.
This study also designed a new type of microprobe apparatus that is low-cost, re-
producible, and easily reparable in case of probe tip breakage. This new probe also
eliminated ow eld disturbances which were observed with the previous probe design
(refer to Chapters 7 and 9). A schematic of the microprobe setup is shown in Figure
4.6. The novelty of the design is the probe tip and the ttings used for connecting it to
the dual-stage pump via heated 1,4

stainless steel tubing. The probe tip is made of a


fused silica tubing typically found in gas chromatography applications. These commer-
cially available tubings are manufactured according to strict inner and outer diameter
specications. This study used a fused silica tubing with an ID of 200jm and an OD of
360jm
18
. It should be noted that the fused silica tubing is manufactured with a poly-
imide resin outer coating which aids in sealing within couplings and provides exibility to
the otherwise brittle silica material. When the probe tip is exposed to high temperatures
in a ame, this coating quickly burns o and does not contaminate the sampling system.
A stainless steel 1,4

-to-1,16

coupling
19
with a 1,16

graphite-reinforced composite re-


ducing ferrule
20
connects the 360jm OD fused silica tubing to the 1,4

stainless steel
sampling line. The entire apparatus is capable of being heated to 350

C.
Extra precautions were taken to eliminate leakage of gases from the surrounding
environment into the sampling line. Leaks into the sampling system dilute the sample
gas and lead to incorrect measurements. For this reason, Swagelok couplings are used for
17
Gauge pressure measured by a vacuum pressure gauge
18
Agilent Deactivated Fused Silica Retention Gap
19
VICI Valco 1/4

to 1/16

external-internal reducing
20
VICI Valco 1/16

one piece fused silica adapter, 0.35 mm tubing OD


Chapter 4. Experimental Apparatus and Analytical Methodology 56
Figure 4.6: Schematic of microprobe (not to scale)
all connections. Leaks into the sampling line were detected using a container lled with
dry ice (i.e solidied carbon dioxide). The container was placed near a suspected point
of leakage, and carbon dioxide gas lls the surrounding area. The NDIR analyzer (see
section 4.6.1) was connected to the sampling line and the dual-stage pump was turned
on. If a spike in CO
2
concentration was observed on the analyzer, then there was a leak
present in the sampling line. The necessary steps were then carried out to eliminate the
leak (e.g., the couplings were changed, the tube was replaced, etc.).
4.5.2 Sampling Procedure
The sampling objective was to obtain samples at various points along the vertical axis
separating the two burner ports. Thus, the sliding stage, on which the microprobe is
mounted, was inserted between the two burner ports. A window was cut into the quartz
shroud enclosing the burner setup to permit insertion of the sampling probe. Figure 4.7 is
a schematic of the probe and burner setup. The tip of the probe was placed approximately
1.5 mm behind the central vertical axis separating the burner ports. This allows samples
to be withdrawn from the middle points of the ame region. After insertion, the probe
was held stationary, while the burner assembly was moved along the vertical axis with
the turn of a micrometer knob on the translation stage. One complete counterclockwise
rotation of the micrometer knob moves the burner assembly downwards by 0.5 mm.
A measuring system is required to dene the exact position of the microprobe between
the two burner ports. For this purpose, the bottom (fuel) port was taken as the zero
Chapter 4. Experimental Apparatus and Analytical Methodology 57
Figure 4.7: Schematic of microprobe and burner setup (not to scale)
height, while the top (oxidizer) port was taken as the maximum height. The position
of the probe was zeroed by touching its tip to the bottom port. The reading on the
micrometer knob was noted as the zero distance. Each counterclockwise turn of the
micrometer knob moved the burner assembly down; thus, increasing the distance between
the probe and the fuel port (i.e., moving the probe upwards). The exact height at which
the probe withdraws samples is equal to the total distance plus the outer radius of the
probe.
Once the probe was positioned at the desired height, the dual-stage pump was turned
on and gases withdrawn from the ame ll the sampling line. The sampling lines were
purged before any analytical measurements were taken. The purge time varied depending
on the location of the probe in the ame. Sampling points away from the center of the
ame required short purge times (e.g., 15-20 minutes), since the low temperature, high
density gases permit high owrates. In contrast, sampling points near the center of the
ame required longer purge times (e.g., 40-45 minutes), since gases in this high temper-
ature region have an extremely low density and permit low owrates. The sampling line
lled with gases from the specied ame region were then ready for analysis.
Chapter 4. Experimental Apparatus and Analytical Methodology 58
4.6 Analytical Techniques
The hydrocarbon species and all oxygenated compounds were analyzed by a GC/FID.
Carbon dioxide and carbon monoxide were quantied by NDIR analysis. The ame
temperature was measured by an R-type thermocouple.
4.6.1 Non-Dispersive Infrared Analysis
NDIR analysis is a technique used to measure gas concentrations based on the energy
absorption characteristics of a gas in the infrared red region. The NDIR instrument
passes infrared light through two identical cells, in parallel, and then onto a detector.
The rst cell is lled with nitrogen, which does not absorb the light and serves as a
reference cell. The second cell contains the sample gas, which absorbs infrared energy.
The detector measures the dierence in energy between the two streams of light. This
dierence is the absorption, which is proportional to the concentration of sample gas by
the Beer-Lambert Law in Equation 4.3:
= c / c (4.3)
where
A is the gas absorbance.
c is the molar extinction coecient (concentration
1
length
1
).
b is the path length that the beam travels in the sampling tube.
c is the gas concentration.
CO and CO
2
Measurements
The NDIR instrument
21
was used to quantify levels of CO and CO
2
in the ame samples.
The instrument is capable of measuring concentrations from 0% to 40%. Initially, the
instrument is zeroed and calibrated. To zero the instrument, nitrogen was owed through
the unit and the unit was zeroed. Next, the unit was calibrated with a gas mixture
containing 9.9% CO and 9.9% CO
2
. The gas mixture was passed through the detector,
and the span was set to match the calibration gas concentration.
Measurements were taken using the dual-stage pump to withdraw samples from the
ame and then pushing them towards the NDIR analyzer. Water, which is damaging
21
NOVA NDIR Analyzer
Chapter 4. Experimental Apparatus and Analytical Methodology 59
to the analyzer, was removed from the sample gas by passing it through a cooling box
and a coalescing lter. The analyzers built-in pump was not used during sampling. The
concentrations were recorded once the displayed reading remains constant for 5 minutes.
As mentioned previously, samples withdrawn from points away from the ame center
have higher owrates than samples from within the middle of the ame. Therefore, the
analyzers display reading stabilized much quicker for the former (e.g., 15-20 minutes)
than the latter (e.g., 40-45 minutes).
4.6.2 Gas Chromatography
Gas chromatography (GC) is a technique used for separating volatile organic compounds
based on their dierences in partitioning between a owing mobile phase and a stationary
phase. In this study, the GC method was used to measure C
1
- C
8
hydrocarbons and
C
1
- C
11
oxygenated species (e.g., esters, aldehydes, alcohols, etc.) A GC instrument
consists of a owing mobile phase (carrier gas), a stationary phase (separation column),
an injection port, an oven, and a detector.
The carrier gas carries the sample gas through the separation column, and compounds
are separated due to partitioning between the two phases. Since partitioning behaviour is
a strong function of temperature, the separation column is placed inside a temperature-
controlled oven. Separation of compounds with a range of boiling points is achieved by
starting at low temperatures and then increasing the temperature until high boiling point
compounds are eluted. The injection port is always maintained at a temperature higher
than the boiling point of the least volatile compound in the mixture.
The amount of time a given component spends in the separation column is called the
retention time. The retention time of a given component remains the same provided the
mobile phase, stationary phase, temperature control, and gas owrates remain constant.
As each component of the separated sample falls onto the detector, a quantitative re-
sponse in the form of a peak is generated. A series of peaks with the retention time on
the x-axis and the detector (e.g., voltage) on the y-axis is called a chromatogram. The
peak retention time is used to identify each compound, and the peak area is used to
determine the quantity of the compound.
The instrument used in these experiments is a Varian 3800 GC
22
with electronic
ow controllers, a 1079 injector, a methanizer, and two ame ionization detector (FID)s.
22
Remotely controlled by a PC using STAR Chromatography Workstation 6.41
Chapter 4. Experimental Apparatus and Analytical Methodology 60
The FID consists of a hydrogen/air ame and a collector plate. As gases ow from the
separation column, the ame burns organic molecules to produce ions. The ions are
attracted to the collector plate, which generates a voltage depending on the quantity of
ions collected. Additional details of the GC setup are discussed in the follow subsections.
GC Carrier Gas
The carrier gas (mobile phase) used for the GC was 99.997 % helium. Hydrocarbon
and oxygen traps were placed before the column to lter out any contaminants from the
carrier gas. The owrate of helium through the separation column was varied depending
on the compounds being studied. Details are available in subsection GC Measurement
Procedures below.
Injection System
The purpose of the injection system is to load the sample gas onto the separation column.
It consists of a sample loop, gas sampling valve (GSV), and an injector. Flame samples
were pushed into the GC sampling loop by the dual-stage pump. After the sample loop
was purged and the sample was ready for analysis, the GSV
23
rotated and delivered the
sample to the GC.
The GSV is a 10 port rotary valve which directs the sample and carrier gases into
the injector. The GSV has two positions; the ll position and the load position. In the
ll position, the sample gas ows through the 0.25 mL sample loop, while helium carrier
gas ows to the injector and separation column. As the GSV turns to the load position,
the sample trapped in the loop comes in-line with the carrier gas ow. The sample is
carried through into the injector and is directed into the column. The GSVs duration
in the load position is a user controlled parameter, which is set to 2 minutes in these
experiments. As the GSV returns to the ll position, the normal ow pattern resumes.
The GC has a 1079 universal capillary injector, which can be run in several modes
based on the type of injector insert used. The injector temperature was set at 250

C to
prevent condensation of sample components. An unpacked 3.4 mm ID insert for splitless
mode operation was used. The sample gas in the injector can be introduced to the column
via split mode or splitless mode. In splitless mode, the entire sample is loaded onto the
column. This mode is advantageous for detecting trace compounds in the sample gas.
23
VICI Valco 10 port valve with air actuator
Chapter 4. Experimental Apparatus and Analytical Methodology 61
However, highly concentrated samples can damage the separation column, so split mode
can be used to load only a portion of the sample onto the column. In these experiments,
a dual column GC method was used, so splitless injection was employed to maximize the
amount of analyte reaching the columns.
GC Measurement Procedures
Two dierent GC measurement procedures have been developed to analyze the gaseous
samples. The rst method was used to study a number of hydrocarbon compounds and
oxygenates in all the ames presented earlier. However, this method was not suitable for
analyzing fatty acid methyl ester compounds because the columns used are not capable
of separating them. Therefore, a second method was developed using a column suitable
for fatty acid methyl ester separation.
The rst method used two separation columns, a methanizer, and two FIDs to study
a number of hydrocarbon and oxygenated compounds in a single run. A schematic of
this method is shown in Figure 4.8 and the specic operating parameters are in Table
4.3. As the sample passes through the injector, it enters a 0.5 meter fused-silica retention
gap
24
. The sample then passes through a y-splitter
25
where it is split to two columns
for separation. C
1
-C
8
hydrocarbons are separated on a non-polar phase HP-Al/S PLOT
capillary column
26
, and then detected on the front FID. Aldehydes, ketones, and alcohols
are separated on a polar phase Poraplot U capillary column
27
, and then passed through
a methanizer before detection.
The FID provides a detector response proportional to the analytes concentration and
number of carbon atoms. For example, 1 mole methane would produce exactly 1,2 the
detector signal of 1 mole ethane, and 1,3 the signal of 1 mol propane. Schoeld [13] has
provided a table of molar response factors for a number of hydrocarbons and oxygenates.
Oxygenated species provide lower responses on an FID because carbon-oxygen bonds are
not broken in the ame. Therefore, species such as formaldehyde and carbon monoxide
provide zero response on an FID, while higher oxygenated hydrocarbons provide weaker
responses than their non-oxygenated counterparts. For example, 1 mole acetaldehyde
has two carbon atoms but produces the same detector signal of 1 mole of methane. In
24
Varian 0.53mm ID retention gap methyl deactivated
25
Varian universal quick seal splitter
26
Agilent Technologies HP-Al/S 50 m x 0.53 mm (L x ID)
27
Varian Poraplot U 25m x 0.53mm (L x ID)
Chapter 4. Experimental Apparatus and Analytical Methodology 62
order to improve the FIDs response to oxygenated hydrocarbons, this study passed the
separated sample gas through a methanizer prior to the FID. The methanizer is a 1,16

stainless steel tube packed with a powdered nickel catalyst. The methanizer is heated to
380

C and requires a constant ow of hydrogen gas. As the sample ows through the
methanizer, the nickel catalyst breaks the carbon-oxygen bonds and then saturates them
with hydrogen. Therefore, all carbon-oxygen bonds are eectively converted to carbon-
hydrogen bonds, and the analyte now appears on the FID as an alkane with the original
number of carbon atoms. For example, formaldehyde appears as methane, acetaldehyde
appears as ethane, etc.
Front
FID
GSV
Sample in
1079
injector
GC column oven
Retention gap
Plot
column
Poraplot U
column
Y-splitter
Rear
FID
methanizer
H
2
Air He
Figure 4.8: Schematic of dual column GC Setup
The second GC method used in these experiments was designed to analyze fatty acid
methyl ester compounds in the methyl decanoate ame using a polar phase DB-WAX col-
umn
28
. This method used only one column and one FID. The GC parameters are provided
in Table 4.4. The method was capable of separating and identifying methyl propanoate,
methyl 2-proepnoate, methyl butanoate, methyl 2-butenoate, methyl 3-butenoate, methyl
pentanoate, methyl 4-pentenoate, methyl hexanoate, methyl 5-hexenoate, methyl hep-
tanoate, methyl 6-heptenoate, methyl octanoate, methyl 7-octenoate, methyl 2-octenoate,
and methyl decanoate.
28
Agilent Technologies DB-WAX 30 m x 0.53 mm (L x ID)
Chapter 4. Experimental Apparatus and Analytical Methodology 63
Table 4.3: Dual Column GC Method Parameters
Front Injector 250

C
2.0 mL/min He ow to column
Splitless injection
Gas Sampling Valve 200

C
Switch to sampling position for 2 min
Column Oven Program 50

C, hold for 3 min
20

C/min to 150

C, hold for 5 min
20

C/min to 180

C, hold for 45 min
Front FID 300

C
30 mL/min H
2
, 300 mL/min air
26 mL/min He makeup
Rear FID 300

C
15 mL/min H
2
, 300 mL/min air
28 mL/min He makeup
Methanizer 380

C
20 mL/min H
2
GC Calibration Procedure
The GC was calibrated for hydrocarbons using four dierent Scotty calibration gas mix-
tures: 100 ppm C
1
-C
6
alkanes in nitrogen; 1000 ppm C
2
-C
6
alkenes in nitrogen; 15 ppm
C
2
-C
4
alkynes in nitrogen; and 100 ppm benzene in air. The calibration was performed
by owing each gas mixture directly into the GC sample loop. The operating condi-
tions for the GC were identical to those mentioned above for the dual column method.
Each calibration gas mixture was used to determine the retention time for each species,
thereby providing qualitative information. The retention times for 1-heptene, 1-octene,
methanol, ethanol, methyl-2-propenoate, methyl-3-butenoate, methyl-4-pentenoate, and
methyl-5-hexenoate were obtained by injecting the saturated vapor above the pure com-
ponent liquid. Quantitative calibration was performed using the 1000 ppm mixture of
ethylene, propene, 1-butene, 1-pentene, and 1-hexene to obtain the FIDs response sig-
nal. The response signal for equivalent concentrations of other hydrocarbons was then
calculated using Schoelds [13] table of molar response factors for hydrocarbon species,
as summarized in Table 4.5. For example, if the FIDs measured response for 1000 ppm
Chapter 4. Experimental Apparatus and Analytical Methodology 64
Table 4.4: GC Method Parameters for Fatty Acid Methyl Esters
Front Injector 250

C
0.5 mL/min He ow to column
Splitless injection
Gas Sampling Valve 200

C
Switch to sampling position for 2 min
Column Oven Program 50

C
10

C/min to 230

C, hold for 5 min
Front FID 300

C
30 mL/min H
2
, 300 mL/min air
26 mL/min He makeup
ethylene was 60,000 counts, then the response for 1000 ppm methane was set at 30,000
counts since the FID carbon number for methane is 1.0 and that of ethylene is 2.0.
GC calibration for the aldehyde and ketone species was dicult because gas cylinders
are not reliable for calibration. These compounds are highly reactive and susceptible to
rapid degradation when stored in gas cylinders. In this study, aldehydes and ketones
species were calibrated using permeation devices. The permeation device is an inert per-
meable tube lled with a pure chemical compound in gas-liquid or gas-solid equilibrium.
When heated to a specied temperature in a passivated glass-coated chamber, the device
emits the compound through the permeable tube wall at a constant mass ow rate (e.g.,
ng/min). An inert carrier gas sweeps over the permeation tube at a constant ow rate
to generate a gas mixture with a known concentration.
Permeation devices were used for calibration of formaldehyde, acetaldehyde, acetone,
propanal, butanal, and acrolein. A schematic of the permeation tube calibration setup
is shown in Figure 4.9. The permeation device was placed within an oven
29
to maintain
the device at a constant temperature. The nitrogen carrier gas was rst puried and then
delivered to the oven chamber using a mass ow controller
30
. A calibration gas mixture
was generated as the carrier gas swept over the permeation device, and this mixture was
sent to the GC. Various calibration concentrations were established by varying the carrier
gas owrate. The concentration of the gas mixture was determined using Equation 4.4.
29
VICI Dynacalibrator Model 150
30
Brooks Mass Flow Controller Model 5850E
Chapter 4. Experimental Apparatus and Analytical Methodology 65
Table 4.5: Measured FID relative molar response factors for organic molecules [13]
Molecule FID relative molar response factor
Methane 1.0
Ethane, Ethylene 2.0
Acetylene 2.2
Propane, Propylene 3.0
Propyne 3.4
Butane, Isobutane 4.0
1-Butene, trans-2-Butene, Cis-2-Butene 4.0
1-Butyne, 2-Butyne, 1,3 Butadiene 4.0
Pentane, 1-Pentene 5.0
Hexane, 1-Hexene 6.0
Benzene 6.0
1-Heptene 7.0
1-Octene 8.0
C =
1 (24.46,:u)
1
c
(4.4)
where
C is the concentration in ppm by mole
mw is the molecular weight of the compound in g/mole
P is the permeation rate in ng/min (provided by the manufacturer)
F
c
is the total ow of the carrier gas in mL/min
24.46 L is the molar volume of nitrogen at STP
GC calibration for the biofuels (i.e., n-butanol, methyl octanoate, and methyl de-
canoate) involved using the fuel delivery and burner setup because gas cylinders or per-
meation tubes were not available. The liquid fuels were pumped into the vaporization and
diluted with nitrogen gas. The exact concentration of the nitrogen-oxygenate mixture
was determined from the nitrogen gas owrate and the liquid ow rate into the mixing
column. This calibration gas then owed to the bottom burner port, and samples were
obtained using the microprobe sampling technique. The aforementioned GC sampling
methods were used to calibrate the system.
Chapter 4. Experimental Apparatus and Analytical Methodology 66
permeation device
N
2
gas purifier
mass flow controller
to GC
oven
90 C
Figure 4.9: Schematic of the permeation tube setup
4.6.3 Temperature Measurement
The ame temperature prole was obtained by measuring the local temperature at var-
ious regions in the ame using an apparatus similar to that described by McEnally et
al. [14]. The measurements were obtained by the most direct method: inserting a ther-
mocouple into the ame. The thermocouple is small compared to the thickness of the
ame front so it does not disturb the ame. The drawbacks of using a thermocouple are
the aerodynamic wake behind the ame front and the possible catalytic activity of the
thermocouple material[10].
The thermocouple measures the temperature by employing the dierence in thermo
electrical properties of dierent metals. When two dissimilar conductors are welded to-
gether and the junction is place in a hot spot, an electric potential is generated which is
proportional to the dierence in temperature between the hot and cold junction. Thermo-
couples made of thin noble metal wires such as platinum and rhodium are advantageous
for the following reasons: they allow a high resolution to be obtained; the aerodynamic
disturbance of the ame front is minimized; and the materials can withstand high tem-
perature environments [10].
Chapter 4. Experimental Apparatus and Analytical Methodology 67
A schematic of the thermocouple apparatus is shown in Figure 4.10. The ame
temperature was measured by inserting the apparatus between the two burners. The
two R-type thermocouple wires
31
(legs) are made of dissimilar metals butt-welded at a
junction. The positive leg is made of pure platinum while the negative leg is comprised of
87% platinum and 13% rhodium. Each wires diameter is 254 jm while the butt-welded
junction is approximately 500 jm in diameter. The wires are housed in ceramic tubes
which provide support while spreading the wires apart. The junction lies between the
ceramic tubes and it is placed where the temperature is measured,. At the opposite end,
the ceramic tubes are attached to an aluminum plate which is mounted on a sliding rail.
One tube is xed in position with a lock nut while the other tube rotates freely upon a
threaded bolt. A spring pulls the free tube towards the xed tube at one to keep the
wires taught at the other end. The positive and negative legs are connected to an R-type
extension wire which carries the measured signal to a digital thermometer
32
.
R-type extension wire
_
+
Thermocouple wires
Fixed ceramic tube
Freely rotating ceramic tube
Thermocouple junction
Spring
Base plate
Digital reader
Figure 4.10: Thermocouple Schematic
31
Omega R-type butt-welded unsheathed ne-gauge thermocouple wires
32
Digisense DualLog R Thermocouple Thermometer
Chapter 4. Experimental Apparatus and Analytical Methodology 68
Correction for Radiation Losses
The temperature measured by the thermocouple diers from the true ame temperature
due to aerodynamic, thermal, and/or chemical perturbations. The methods of minimizing
the eect of these perturbations are discussed in detail by Fristrom and Westenberg [6].
However, even if all disturbances are minimized, the thermocouple will register a dierent
temperature than the true stream temperature due to radiation losses. Correcting for
these losses is estimated by equating the heat transferred to the thermocouple from the
gas to the heat lost by radiation from the wires. The equation given for a spherical device
(i.e., Nusselt number of 2) is:
1
j
1
c
=
c o d (1
4
c
1
4
&
)
2 /
(4.5)
where
T
j
is the true gas temperature (K)
T
c
is the measured gas temperature (K)
c is the emissivity of the thermocouple element (dimensionless)
o is the Stefan-Boltzmann Constant = 5.67x10
08
(
W
n
2
1
4
)
d is the wire diameter (m)
k is the thermal conductivity of gas (
W
n1
)
T
&
is the wall (ambient) temp. to which heat is radiated = 300 (K)
The thermal conductivity, k, of the gases was estimated as the thermal conductivity of
air at the measured temperature. The thermal conductivity of air at various temperatures
was obtained from the CRC Handbook [3]. Linear interpolation and extrapolation was
used to determine thermal conductivity at temperatures not listed in the handbook.
The emissivity, c, of the thermocouple element was obtained from the study by
Bradley and Entwistle [15].
Literature Cited
[1] S. Sarathy, Using an opposed ow diusion ame to study the oxidation of C4 fatty
acid methyl esters, Masters thesis, University of Toronto, 2006.
[2] I. Glassman, Combustion, 3rd ed. San Diego, CA: Academic Press, 1996.
[3] D. R. Lide, Ed., CRC Handbook of Chemistry and Physics, 87th Edition. Boca Ra-
ton, FL: Taylor and Francis, 2007. [Online]. Available: http:/www.hbcpnetbase.com
[4] K. Seshadri, T. Lu, O. Herbinet, S. B. Humer, U. Niemann, W. J. Pitz, R. Seiser,
and C. K. Law, Experimental and kinetic modeling study of extinction and ignition
of methyl decanoate in laminar non-premixed ows, Proceedings of the Combustion
Institute, vol. 32, no. Part 1, pp. 10671074, 2009.
[5] S. M. Schoenung and R. K. Hanson, CO and temperature measurements in a at
ame by laser absorption spectroscopy and probe techniques. Combustion Science
and Technology, vol. 24, no. 5-6, pp. 227 237, 1981.
[6] R. M. Fristrom, Comments on quenching mechanisms in the microprobe sampling
of ames, Combustion and Flame, vol. 50, pp. 239242, 1983.
[7] A. M. Vincitore and S. M. Senkan, Polycyclic aromatic hydrocarbon formation in
opposed ow diusion ames of ethane, Combustion and Flame, vol. 114, no. 1-2,
pp. 259266, July 1998.
[8] A. Sinha and M. J. Thomson, The chemical structures of opposed ow diu-
sion ames of c3 oxygenated hydrocarbons (isopropanol, dimethoxy methane, and
dimethyl carbonate) and their mixtures, Combustion and Flame, vol. 136, no. 4,
pp. 548556, March 2004.
69
Literature Cited 70
[9] S. S. M. Kassem, M. Qun, Chemical structure of fuel-rich 1,2-
C2H4Cl2/CH4/O2/Ar ames: Eects of micro-probe cooling on the sampling
of ames of chlorinated hydrocarbons, Combustion Science and Technology,
vol. 67, pp. 147157, 1989.
[10] R. M. Fristrom, Flame structure. New York: McGraw-Hill, 1965.
[11] U. Struckmeier, P. Osswald, T. Kasper, L. Boehling, M. Heusing, M. Koehler,
A. Brockhinke, and K. Kohse-Hoeinghaus, Sampling Probe Inuences on Temper-
ature and Species Concentrations in Molecular Beam Mass Spectroscopic Investiga-
tions of Flat Premixed Low-pressure Flames, Zeitschrift Fur Physikalische Chemie-
International Journal of Research in Physical Chemistry & Chemical Physics, vol.
223, no. 4-5, pp. 503537, 2009.
[12] S. A. Syed, Oxidation studies of surrogate bio-diesel fuels in opposed ow diusion
ames, 2005.
[13] K. Schoeld, The enigmatic mechanism of the ame ionization detector: Its over-
looked implications for fossil fuel combustion modeling, Progress in Energy and
Combustion Science, vol. 34, no. 3, pp. 330350, June 2008.
[14] C. McEnally, U. Koylu, L. Pfeerle, and D. Rosner, Soot volume fraction and
temperature measurements in laminar nonpremixed ames using thermocouples,
Combustion and Flame, vol. 109, no. 4, pp. 701720, June 1997.
[15] D. Bradley and A. Entwistle, Deterimination of the emissivity, for total radiation,
of small diameter platinum-10% rhodium wires in the temperature range 600-450
degrees C, British Journal of Applied Physics, vol. 12, pp. 708711, December 1961.
Part II
Biobutanol
71
Chapter 5
Background
In 2006, British Petroleum (BP) and DuPont announced that they would start selling
biobutanol made from sugar beets, as a gasoline blending component in the United
Kingdom [1]. The biobutanol, would be produced using a fermentation process similar
to that of ethanol, and its feedstock include sugar beet, sugar cane, corn, wheat, and
potentially lignocellulosic biomass. Proponents [2, 3, 4] of biobutanol highlight many
characteristics that make it a superior biofuel, such as:
the ability for blending with gasoline at higher concentrations than ethanol without
modifying current vehicle and engine technologies;
a higher energy density than ethanol, which provides better fuel economy at higher
blend ratios;
enhanced water tolerance compared to ethanol, allowing use of the existing fuel
distribution infrastructure;
the use of existing feedstock and reneries (with minor modications) used for
bioethanol production;
the potential to produce from lignocellulosic and waste biomass feedstock;
and a lower vapour pressure than ethanol and gasoline, thereby decreasing volatile
organic compounds (VOC) emissions.
It is apparent that biobutanol is being proposed as an eventual replacement for the
bioethanol currently used in SI engines. BP and DuPont stated that a complete environ-
mental LCA based on actual manufacturing design models is underway, but to date no
72
Chapter 5. Background 73
results have been published [5]. If biobutanol is to replace bioethanol, then its sustain-
ability needs to be critically assessed. In addition, the combustion kinetics of biobutanol
must be understood to aid in the design of combustion systems. This chapter provides
background information relevant to the development of an LCA and a chemical kinetic
mechanism for biobutanol.
5.1 Biobutanol History
The production of biobutanol from biomass feedstock is not novel. Jones and Woods [6]
have provided a detailed history of biobutanol production via fermentation. Figure 5.1
is a visual representation of this history and following is a summary.
Figure 5.1: Timeline of biobutanol history
The fermentation of sugars to biobutanol, specically the isomer n-butanol
1
, was rst
documented by Louis Pasteur in 1861, although the yields he achieved were not sucient
for commercialization. In the early 1900s, a German born chemist, Chaim Weizmann, was
attempting to produce butanol for synthetic rubber production, and discovered that the
Clostridium acetobutylicum organism was capable of converting large amounts of sugars
into a mixture of acetone-butanol-ethanol (ABE) in the molar ratio of 3:6:1. Weizmann
1
Henceforth, the terms biobutanol, n-butanol, and butanol are used interchangeably
Chapter 5. Background 74
had performed his pioneering work while living in England, and at the onset of World
War 1, the English sought his help in producing bioacetone from biomass as a precursor
to cordite, a smokeless propellant used in munition. During the war, a large amount of
ABE was produced in England, Canada, and the U.S. from feedstock such as corn and
sugar beet molasses. During this period, the unwanted butanol was stored in large vats
waiting for some commercial use.
After World War 1, the automotive industry in the U.S. was rapidly growing and
the demand for paint lacquers increased dramatically. One company commercialized a
method of producing high quality lacquers using butanol and its ester, butyl acetate, as
solvents. Thus, in the 1920s and 1930s, there was a large industry built around convert-
ing biomass sugars into ABE for the automotive paints industry. In a 1927 Scientic
American article, Killeer [7] states that the butanol production in the U.S. doubled to
60 tons per day in just 18 months.
In 1938, World War 2 began and all the U.S. fermentation facilities reverted to pro-
ducing acetone for cordite manufacturing. Japan, South Africa, India, Australia, China
and the U.S.S.R also opened a number of ABE fermentation facilities using wheat, rye,
and corn as feedstock. After World War 2, there was a rapid decline in ABE production
from biomass for two reasons. The cost of biomass feedstock rose sharply as corn, wheat,
and molasses became staples for animal feed. In addition, the petrochemical industry
boomed in the 1950s and 1960s, leading to superior solvents for use in automotive paint
lacquers. The demand for butanol decreased sharply, and by the 1980s virtually all the
ABE plants in the world had closed.
5.2 Biobutanol Production
2
Butanol can be produced from the same feedstock as ethanol (e.g., sugar beet, sugar
cane, corn, wheat, and lignocellulosic biomass) using a fermentation process. During
the rst part of the last century, the production of butanol was a large-scale industrial
process. The conventional production process used C. acetobutylicum to produce acetone,
n-butanol, and ethanol (termed the ABE process). Almost two thirds of the total butanol
produced in the late 1940s was produced by fermentation. The traditional fermentation
process could not compete economically with the production of butanol from petroleum
2
This sections includes background research conducted by S.M. Sarathy and Y. Zhang (PhD Candi-
date, Dept. of Civil Engineering, University of Toronto).
Chapter 5. Background 75
due primarily to high feedstock (e.g., corn) costs and the large energy requirements for
butanol recovery [6].
The traditional process for production of butanol and ethanol from corn is as follows,
[8, 9]:
1. The corn crop is cultivated and milled.
2. The milled corn is slurried with water. For ethanol, the milled corn undergoes liq-
uefaction and sacharrication to convert starches to monosaccharides. This process
is not required for butanol because the fermentation organism is capable converting
starches to ABE.
3. The slurried mixture is then fermented in a batch reactor to convert sugars (and
starches) into a beer solution containing the desired product(s).
4. The desired solvent, or biofuel, is then recovered from the beer by a series of
separation processes.
The major consumer of energy in the production of butanol and ethanol are the
separation processes (i.e., distillation) required for product recovery. The production of
butanol via fermentation is severely handicapped due to the low concentrations of butanol
in the fermented beer. The C. acetobutylicum used to convert starches and sugars into
ABE cannot tolerate high solvent concentrations, so typical end-product concentration
is 20 g of ABE per liter of beer [8].
Several recent advances have occurred in butanol production, including the develop-
ment of strain, C. beijerinckii BA 101, with increased solvent tolerance to 33g/L total
solvents [10, 11], and the development of advanced fermentation techniques and down-
stream product recovery processes [12]. Environmental Energy Inc. (EEI), now Butyl-
Fuels LLC, has patented a novel dual-stage process and claimed the process signicantly
improves butanol yield and minimizes undesired byproducts [3]. Green Biologics, a com-
pany based in the United Kingdom, has obtained signicant funding to commercialize
the conventional ABE fermentation process by utilizing waste feedstock and advanced
fermentation and separation processes [4].
Historically, starch and sugar feedstock have been utilized for butanol production,
however, butanol could be produced from lignocellulosic biomass, a more abundant feed-
stock. During the last several decades, research has been conducted to convert lignocel-
lulosic biomass to butanol. Yu et al. [13] report that C. acetobutylicum can convert both
Chapter 5. Background 76
hexose and pentose sugars present in biomass hydrolyzates to butanol. Recently, Zverlov
et al. [14] documented the production of butanol from hydrolyzates of lignocellulosic
waste at a full-scale industrial plant in the 1980s in Russia.
Current metabolic engineering research is attempting to resolve the end-product in-
hibition associated with the conventional ABE process, so that higher yields of butanol
can be achieved. BP and DuPont disclosed that their partnership is patenting novel
biocatalysts for high yield production of n-butanol, as well as its higher octane isomers,
sec-butanol and iso-butanol [5]. Atsumi et al. are genetically modifying the metabolic
pathways of E. Coli to develop non-fermentative pathways for biofuel n-butanol, sec-
butanol, and iso-butanol production [15].
The aforementioned genetic engineering research may lead to novel biobutanol pro-
duction technologies in the long term (e.g., 15-20 years); however, in the short term (e.g.,
5-10 years), biobutanol could only be produced at industrial scales by reintroducing the
traditional fermentation technology, albeit with process improvements. Presently, there
is no commercial-scale production of biobutanol from starch, sugar or lignocellulosic feed-
stock and there is limited data on the processes. In spite of this, there is renewed interest
in butanol, particularly considering its apparent attractiveness as an automotive fuel.
5.3 Biobutanol Fuel Properties
3
Although limited research has been conducted on the use of butanol in vehicles, re-
cent testing by BP found that butanol has several fuel property advantages compared
to ethanol [16]. Selected fuel properties are presented in Table 5.1. Butanol has a
higher energy density (i.e., LHV) than ethanol, which leads to better vehicle fuel econ-
omy. While ethanol has a higher octane rating than gasoline, butanol is octane neutral,
thus lessening requirements for additional fuel modication when blending with gasoline.
The lower oxygen content of butanol allows it to be blended at higher proportions than
ethanol without requiring engine modications or exceeding current blending regulations.
When blended with gasoline, butanol contributes much less to the reid vapour pressure
equivalent (RVP) than ethanol, even having a negative impact on the RVP when bu-
tanol and ethanol are co-blended. Although ethanol/gasoline blends were found to lead
to distillation curve abnormalities, which may negatively impact driveability, this was
3
This sections includes background research conducted by S.M. Sarathy and S. Sleep (Undergraduate
Research Assistant, Dept. of Civil Engineering, University of Toronto).
Chapter 5. Background 77
Table 5.1: Selected fuel properties of butanol, ethanol, and gasoline
n-Butanol Ethanol Gasoline
Chemical Formula C
4
H
9
OH C
2
H
5
OH C
4
to C
12
hydrocarbons
Energy Density (LHV)(MJ/L) 26.9 21.2 32.2-32.9
Fuel Density at 20

C (kg/L) 0.81 0.79 0.72-075
Boiling Point (

C) 117 78 <210
Octane Number (R+M/2)
a
87 116 87
RVP at 10% v/v, (kPa) 34 130 <60/90
b
Oxygen (%wt) 21.6 34.7 <2.7
a
Octane number of alcohols is given for blends containing 5 vol% alcohol in
gasoline.
b
Summer/Winter specications provided for gasoline DVPE.
not the case with butanol/gasoline blends, as no vapour pressure or distillation curve
abnormalities were noted. Butanol, unlike ethanol, is immiscible with water, so it can
be distributed in existing pipelines without the risk of water contamination; ethanol re-
quires an alternate distribution infrastructure. In addition, butanol is not corrosive to
engine components while ethanol corrodes copper and brass. Preliminary results of BPs
vehicle testing of gasoline/butanol blends containing 5% and 10% butanol showed no
signicant changes in carbon monoxide, hydrocarbon, and NOx emissions compared to
regular gasoline and gasoline/ethanol blends containing 5% ethanol. These combined fac-
tors indicates that butanol may be a superior gasoline blending component than ethanol.
However, further combustion testing and evaluation of the life cycle performance of bu-
tanol compared with ethanol and gasoline are required.
Literature Cited
[1] G. Hess, BP and DuPont to make biobutanol, Chemical & Engineering News,
vol. 84, pp. 910, 2006.
[2] BP-DuPont, Bp and duPont biofuels fact sheet, British Petroluem and DuPont,
Tech. Rep., June 2006.
[3] D. Ramey and S. Yang, Production of butyric acid and butanol from biomass,
US Department of Energy, Morgantown, Washington, Tech. Rep. DE-F-G02-
00ER86106, 2004.
[4] G. Clark, Green biologics secures 1.58 million British pounds to develop next gen-
eration biofuel, Biofuel Review, vol. October, 2007.
[5] BP and DuPont, DuPont and BP disclose advanced biofuels partnership targeting
multiple butanol molecules, BP America Press Release, 2008.
[6] D. Jones and D. Woods, Acetone-butanol fermentation revisited, Microbiological
Reviews, vol. 50, no. 4, pp. 484524, December 1986.
[7] D. Killeer, A microbe in international aairs, Scientic American, 1927.
[8] N. Qureshi and H. Blaschek, Economics of butanol fermentation using hyper-
butanol producing clostridium beijerinckii ba101, The Institution of Chemical En-
gineers, vol. 78, pp. 139144, 2000.
[9] J. R. Kwiatkowski, A. J. McAloon, F. Taylor, and D. B. Johnston, Modeling the
process and costs of fuel ethanol production by the corn dry-grind process, Indus-
trial Crops and Products, pp. 288296, 2006.
78
Literature Cited 79
[10] T. Ezeji, N. Qureshi, and H. Blaschek, Production of acetone, butanol and ethanol
by clostridium beijerinckii ba101 and in situ recovery by gas stripping, World Jour-
nal of Microbiology & Biotechnology, vol. 19, pp. 595603, 2003.
[11] N. Qureshi and H. Blaschek, Evaluation of recent advances in butanol fermentation,
upstream, and downstream processions. Bioprocess and Biosystems Engineering,
vol. 24, pp. 219226, 2001.
[12] T. Ezeji, N. Qureshi, and H. Blaschek, Bioproduction of butanol from biomass:
from genes to bioreactors. Current Opinion in Biotechnology, vol. 18, pp. 220227,
2007.
[13] E. K. C. Yu, L. Deschatelets, and J. Saddler, The bioconversion of wood hy-
drolyzates to butanol and butanediol, Biotechnology Letters, vol. 6, pp. 327332,
1984.
[14] V. Zverlov, O. Berezina, G. Velikodvorskaya, and W. Schwarz, Bacterial acetone
and butanol production by industrial fermentation in the soviet union: use of hy-
drolyzed agricultural waste for biorenery, Applied Microbiology Biotechnology,
vol. 71, pp. 687597, 2006.
[15] S. Atsumi, T. Hanai, and J. Liao, Non-fermentative pathways for synthesis of
branched-chain higher alcohols as biofuels, Nature, vol. 451, pp. 8689, 2008.
[16] L. Wolf. (2007, March) 1-butanol as a gasoline blending biocomponent.
U.S. Environmental Protection Agency Clean Air Act Advisory Com-
mittee Mobile Sources Technical Review Subcommittee. [Online]. Available:
http://www.epa.gov/air/caaac/mstrs/March2007/Wolf.pdf
Chapter 6
LCA of Biobutanol for use in
Transportation
6.1 Introduction
Ethanol from sugarcane and corn, and biodiesel from soybean and rapeseed are being
commercially produced today; however, the biofuels of choice for the future and their
methods of production are still uncertain [1, 2]. Initiatives to substantially increase bio-
fuel production and upcoming low carbon fuel standards motivate a critical examination
of the environmental implications of current and potential future fuel production and
end use pathways.
Over the past two decades, there have been many studies examining the environmen-
tal performance of corn-based ethanol. However, few studies have been conducted on
corn-based butanol. It is only recently that butanol production technology has become
competitive due to improved product yields. The following subsections include a review
of important studies on bioethanol LCA, as well as some recently published LCA research
on biobutanol.
6.1.1 Bioethanol LCA
The sustainability of bioethanol has been under intense scrutiny by the scientic com-
munity. Pimentel and Patzek [3, 4] concluded that the total energy required to produce
corn-based ethanol is greater than the energy content of the ethanol produced. However,
the majority of studies [5, 6, 7, 8, 9, 10] indicate that corn-based ethanol can displace
80
Chapter 6. LCA of Biobutanol for use in Transportation 81
fossil energy sources and reduce GHG emissions. A study by Farrell et al. developed
a meta-model to evaluate various LCA studies [11] and demystify the aforementioned
discrepancies. The results indicate that Pimentel and Patzek [3, 4] used obsolete data
from ethanol production technologies in the early 1980s, and the authors did not cor-
rectly allocate credits for the ethanol coproducts. Farrell et al. also made the following
recommendations for future LCA evaluations:
use the displacement method to credit coproducts,
obtain accurate and current data,
clearly dene future scenarios, and
dene performance metrics such as GHG emissions, fossil energy use, soil erosion,
etc.
Kim and Dale [8, 9] used the system expansion approach to credit coproducts and also
considered the carbon sequestration eect from increasing the soil organic carbon level
during biomass production. The system boundaries include upstream processes (e.g.,
fertilizer production and transport), ethanol production and coproducts by dry milling,
urea production, and the corn farming system. The study concludes that using ethanol
can reduce GHG emissions and fossil energy use.
Studies at the Argonne National Laboratory (ANL) [10, 12] used the Greenhouse
Gases, Regulated Emissions, and Energy Use in Transportation (GREET) model to
conclude that ethanol could reduce GHG emissions and fossil energy use. The authors
used the displacement method to assign credits to the fuel ethanol coproducts. The
results by Wu et al. [10] project that corn ethanol produced by the dry milling process in
the year 2012 reduces fossil energy use by 38% and GHG emissions by 21%. The GREET
model version 1.8 by ANL is an LCA model with detailed descriptions of all the data sets
and procedures. Therefore, it will be used in this study to determine the environmental
performance of biobutanol and to compare it with bioethanol.
It is clear from the literature that conventional ethanol production from food crops,
most notably corn, has limited potential [13, 14]. Reasons for this include competition
from the food industry [15], limited agricultural land for crop growth [13], and high
energy input requirements for agricultural chemicals and harvesting [14]. For biofuels
to live up to their potential as attractive alternative fuels, lignocellulosic biomass must
Chapter 6. LCA of Biobutanol for use in Transportation 82
be considered. Consisting of energy crops, agricultural and forest residue as well as
other sources, lignocellulosic biomass oers advantages over food crop feedstock. These
advantages include: the ability to be cultivated on marginal agricultural land, lower
agricultural chemical requirements, lower energy requirements, and the potential of uti-
lizing the lignin portion of the biomass as an energy source for the production process.
For these reasons the net energy gains of lignocellulosic ethanol are much higher than
those of corn ethanol, and the resulting GHG emissions are much lower. Despite these
advantages, the conversion of lignocellulosic feedstock to fuels is not yet at commercial
scale, primarily because lignocellulosic biomass is more challenging to convert to alcohols
[13, 14, 15, 16].
Land Use Change and Food Security Issues
Recently, there have been a number of studies citing food security and land use change is-
sues associated with biofuel production. Prices of food, including corn and rice, increased
sharply in 2007 and 2008. Due to media sensationalisation, the rise in food prices was
largely attributed to the increasing diversion of food crops for biofuel use. In reality, the
rise in food prices can be attributed to a number of factors, including, increasing energy
costs, climate change, stagnation in crop productivity, and diversion of crops or crop-
lands to biofuel production [17]. Furthermore, rising food prices may actually benet
the largest population of undernourished individuals who are farmers [18]; however, the
urban poor would suer.
As the value of biofuels and the price of their feedstock (i.e., corn) increase, farmers
worldwide are converting forests and grassland to cropland for additional grain produc-
tion. Searchinger et al. and Fargione et al. report [19, 20] that the conversion of carbon-
rich forest lands to cropland inevitably release CO
2
emissions into the atmosphere. The
magnitude of the GHG released is much larger than the GHG reductions the biofuels oer
by displacing fossil fuels, suggesting that biofuels actually increase GHG emissions via
indirect land use change. Proponents of biofuels argue that the aforementioned studies
overestimate the carbon debt associated with biofuels. In any case, the use of biofuels
derived from traditional crops and croplands is likely unsustainable in the long term,
and therefore, it is vital that biofuels be produced from waste feedstock and/or feedstock
grown on degraded or abandoned land.
Chapter 6. LCA of Biobutanol for use in Transportation 83
6.1.2 Biobutanol LCA
For all the research on ethanol, very little has been published on the life cycle implications
of butanol. S&T Squared Consultants Inc., a consulting company performed an LCA of
biobutanol using the GHGenius model[21]. During the course of the present study, Wu
et al. published results using the GREET model [22]. A comparison of the two studies
shows some inconsistent results primarily due to Wu et al. utilizing recent data [23, 24]
while S&T Squared Consultants Inc. utilized data for the same process (ABE) but from
the mid-1980s. Although both studies found that the butanol production process was
more energy intensive than that of ethanol, they diered substantially in the energy
inputs estimated for the butanol production. S&T Squared Consultants Inc.Squared
Consultants Inc. found that butanol production consumed more fossil energy than the
energy the butanol contained. Wu et al. concluded that butanol resulted in a net
energy gain with GHG emissions slightly higher than ethanol, but still 20% to 60% lower
than gasoline. They also found that the butanol benets were very dependent on the
treatment of the acetone coproduct due to the coproduct credit scheme. It should be
noted that neither study considered lignocellulosic biomass feedstock, or the possible
benet of using existing gasoline pipeline infrastructure for butanol distribution, as both
assumed delivery by diesel truck.
The objective of this study is to examine the potential attractiveness of butanol as
a transportation fuel to replace gasoline use in the U.S. Selected life cycle environmen-
tal metrics associated with the production of butanol from corn grains are compared
with ethanol using the same feedstock. The goal is to determine the fossil energy use,
petroleum energy use, and GHG emissions associated with the production and use of
corn-based butanol. The results of this study are also compared with those of a previous
study [22].
6.2 Methodology
The LCA [25] method is utilized to evaluate the environmental performance of butanol
and ethanol derived from corn in the U.S. The environmental metrics examined are fossil
and petroleum energy use and GHG emissions. The GHGs considered are CO
2
, CH
4
,
and N
2
O which are aggregated as CO
2
equivalent (CO
2
-eq.) based on the 100-year global
warming potentials recommended by [26].
Chapter 6. LCA of Biobutanol for use in Transportation 84
The butanol and ethanol LCA models consist of similar activities. The life cycle
begins with corn farming and the associated manufacture of upstream agronomic inputs
(i.e., fertilizers and herbicides). The corn is collected from the eld and transported to a
conversion facility where it is converted to ethanol or butanol with associated coproducts.
The alcohol is then distributed to a facility for blending with gasoline. The life cycle of
the biofuel ends with the combustion of the alcohol/gasoline blend in an automobile.
Transportation between the various fuel production processes is also included in the
models. Figure 6.1 is a simplied diagram showing the life cycle of butanol, including its
use in a light duty vehicle (LDV).
This study divides the lifecycle into well-to-pump (WTP) and pump-to-wheel (PTW)
portions. The WTP portion includes the life cycle results solely associated with the
production and distribution of neat alcohol fuels, so as to distinguish from those when
gasoline is blended with them. The WTP results for neat alcohol are reported with
the functional unit (FU) being 1 MJ of alcohol available at the pump, where the pump
is dened as the gasoline blending facility. The PTW portion includes blending of the
biofuel with gasoline and subsequent processes, and this study reports the total well-to-
wheel (WTW) results for 85% alcohol and 15% gasoline blends. The FU for the WTW
analysis is one km driven by a midsize passenger vehicle.
6.2.1 Data sources and uncertainties
The often-cited, publicly available GREET 1.8b model contains corn ethanol and butanol
pathways [12]. The mix of energy sources (e.g., coal, natural gas, electricity, etc.) used
for biofuel production are considered the same for both ethanol and butanol, and the
values are the default presented in GREET 1.b. The comparison ethanol pathways are
obtained directly from this model and default values for the year 2010 are assumed. The
year 2010 was selected for the analysis as it is expected that butanol could be available
by that time based on the recent announcements. Models for butanol production are
developed in this paper based on production data available in the research literature, as
described below. The production data are based on experimental work which has not
been optimized from a process engineering perspective. There is considerable uncertainty
as to the feasibility and performance of the process at commercial scales.
Chapter 6. LCA of Biobutanol for use in Transportation 85
Figure 6.1: A simplied life cycle owchart for corn-derived butanol and ethanol
6.2.2 Corn ethanol and butanol production
Corn production and its life cycle implications are identical whether the ultimate fuel
to be produced is ethanol or butanol. Activities included in corn production are: agri-
cultural chemical manufacture, including transport to the eld and application; grain
harvest; and transport to the biofuel production facility. Corn ethanol plants can be
classied into dry milling and wet milling operations. The GREET 1.8b dry mill facility
pathway is utilized since all recent facilities constructed in the U.S. have been of this type.
The resulting ethanol yield is 405 L/Mg of corn with dried grains with solubles (DDGS)
produced as a co-product at a yield of 290 bone-dry kg/Mg of corn [12]. Prior to leaving
the facility, the ethanol is usually denatured with at least 1-5 vol% gasoline; however, in
this study the denaturant is excluded, in order to compare the life cycles of neat biofuels
leaving the ethanol and butanol facilities.
The production of ethanol and butanol from corn using the dry milling process is quite
similar, although as noted earlier ethanol production is a mature process while butanol is
not currently produced at commercial scale. For the butanol life cycle model, all stages
of the life cycle up to the corn delivery to the production facility are assumed identical
Chapter 6. LCA of Biobutanol for use in Transportation 86
to those modeled in the ethanol process. After delivery to the production facility, the
two models diverge as the saccharication, fermentation and separation processes dier.
The energy required for grain handling and saccharication in the butanol facility was
estimated using the USDA AspenPlus (Aspen Technologies, Inc., Cambridge, MA) model
for corn dry mill ethanol [27, 28]. The AspenPlus model was modied for the butanol
process to produce a fermentation broth containing 60 g/L of sugars and assuming 95%
of substrate utilization
1
. The conversion of glucose to acetone, butanol, and ethanol was
estimated based on lab-scale yields using the C. beijerinckii BA101 organism in a batch
reactor [29]. Estimates were based on batch conversion technology since it is widely used
in industrial fermentation technology, and other newer processes based on continuous
fermentation technologies, in-situ gas stripping, and fed-batch technologies [23, 24, 30]
are not.
The energy requirements for the downstream processes (i.e., separation of solvents
from broth) were estimated using an AspenPlus model consisting of a gas stripper and
two distillation columns based on previous work by Liu [31]. The model does not employ
advanced separation technologies, such as adsorption and pervaporation, which have
been proven to reduce energy consumption in recent laboratory experiments [32]. The
downstream process model was not optimized for energy eciency, so it was estimated
that 40% of the energy prediction for the the separation processes could be saved using
heat-integrated distillation [33, 34]. It should be noted that since the completion of the
present study, an AspenPlus simulation of a complete process for producing butanol via
acetone, butanol, and ethanol corn fermentation has been published by Liu [35]. It is
expected that this new simulation could predict energy use better than the simulations
mentioned above since it includes all processes from grain processing through to product
purication in one model.
Beside DDGS, the butanol fermentation process also produces ethanol and acetone
as co-products. Data on butanol and co-product yields are presented in Table 6.1
2
. Co-
product allocation methods could signicantly impact the energy and emissions results
for corn ethanol and butanol. Various scenarios for butanol production were investigated
using dierent co-product allocation methods. The butanol scenarios developed in this
1
These conversion estimates were determined by Dr. M. Grin (Dept. of Civil and Environmental
Engineering, Carnegie Mellon University).
2
The yield and energy use estimates were estimated by S.M. Sarathy and Y. Zhang (Dept. of Civil
Engineering, University of Toronto).
Chapter 6. LCA of Biobutanol for use in Transportation 87
study are as follows
3
Butanol with no co-products
Butanol with DDGS co-product credited by the displacement method (assumes
DDGS displaces corn and soybean meal as an animal feed)
Butanol with DDGS, acetone, and ethanol co-products credited by the energy allo-
cation method (i.e., energy use and associated GHG emissions are allocated among
products based on their energy output shares)
In addition, our butanol scenarios are compared to those of Wu et al. [22] and ethanol
results in GREET 1.8b. The comparison scenarios are as follows:
Wu et al. [22] results for butanol with DDGS and ethanol co-products credited by
the displacement method
Wu et al. [22] results for butanol with DDGS, acetone, and ethanol co-products
credited by the energy allocation method
Ethanol with no co-product (GREET 1.8b) [12]
Ethanol with DDGS co-product credited by the displacement method (GREET
1.8b) [12]
6.2.3 Post Production Life Cycle Activities
Following the production of the biofuels they are transported from the plants to bulk
terminals, where they are blended with gasoline. Barge, rail and truck are assumed to
transport 40%, 40%, and 20% of the ethanol, respectively [12]. The distances for the
barge, rail and truck transport are 837 km, 1287 km, and 129 km, respectively. For
butanol, it is assumed that 100% pipeline transportation with an average distance of 966
km (i.e., the weighted average of the transportation distances for ethanol). 85% of each
biofuel is blended with 15% conventional gasoline (CG) (i.e., not reformulated) before
use in the vehicle to produce 85% ethanol-15% gasoline (E85) and 85% butanol-15%
gasoline (Bu85) blends. After blending, the fuel is distributed to refueling stations by
3
The scenarios presented in this study were selected by Dr. H. MacLean (Dept. of Civil Engineering,
University of Toronto.)
Chapter 6. LCA of Biobutanol for use in Transportation 88
Table 6.1: Data for estimating energy use and GHG emissions for corn butanol production
Assumptions Data Source
Corn Production 2010 Assumptions GREET 1.8b
in GREET 1.8b
Butanol production
Butanol yield (L/Mg of corn) 200 Estimated here
Acetone yield (L/Mg of corn) 100 Estimated here
Ethanol yield (L/Mg of corn) 34 Estimated here
DDGS yield (bone-dry kg/Mg of corn) 280 GREET 1.8b
Process fuel requirement (MJ/L of butanol)
a
40 Estimated here
a
Process fuel requirements are assumed to be met 80% by natural gas and 20% by
coal based on GREET 1.8bs corn ethanol pathway [12].
diesel truck (48 km). The GREET 1.8b model is used to estimate energy use and GHG
emissions of the aforementioned distribution processes.
For the PTW analysis, the fuels are assumed to be utilized in a recent model year,
exible fuel vehicle (FFV), the 2006 Chevrolet Impala. Flexible fuel vehicles can utilize
100% gasoline or an alcohol/gasoline blend with up to 85% alcohol. It is assumed that
there would be only minor modications needed for the vehicle to utilize butanol blends
and that these would not impact vehicle performance with respect to energy use and GHG
emissions. The vehicle operation stage of the life cycle is based on fuel economy values
published by the U.S. Department of Energy [36]. The combined fuel consumption (fuel
economy) (55% city and 45% highway driving) for the Impala when fueled with gasoline
is 9.2 L/100km (25.5 mpg), whereas it is 12.3 L/100km (19.2 mpg) when fueled with E85.
Ethanols lower energy density accounts for the decrease in fuel economy; though, this is
very slightly oset by the greater thermal eciency associated with ethanol combustion
[37]. Butanols higher energy density results in a smaller decrease in volumetric fuel econ-
omy than with ethanol. However, the combustion eciency improvement possible with
ethanol blends does not occur for butanol [37]. Therefore, the calculated fuel economy
for Bu85 is 10.7 L/100km (21.8 mpg), assuming the vehicle eciency is the same as when
fueled with gasoline. When inputted into the GREET model, the fuel economy values
are convereted to gasoline-equivalent units by multiplying the fuel economy by the ratio
of LHV gasoline/LHV alcohol-gasoline blend.
Chapter 6. LCA of Biobutanol for use in Transportation 89
6.3 Results and Discussion
In this section, the life cycle results for the butanol production scenarios are presented
and compared to the model and literature results for butanol and ethanol.
6.3.1 WTP fossil energy use
Figure 6.2
4
presents the WTP fossil energy use for the various scenarios. The results for
scenarios with no co-product(s) are broken down into, corn farming (consisting of the
activities associated with feedstock production), production (conversion of the feedstock
to alcohol) and transportation (distribution of the fuel to blending terminals). Butanol
has a high fossil energy input, as can be seen from the bar showing butanol with no co-
products. The inputs are more than double the inputs of ethanol under similar conditions.
The much higher fossil energy is attributed to the lower yields and higher downstream
product recovery energy requirements in butanol production. However, when the DDGS,
acetone, and ethanol co-products are accounted for using the energy allocation method,
the fossil energy used for butanol production is substantially lower. The results of this
study are higher than those of Wu et al. [22] when no co-products are considered due to
the more conservative butanol yields assumed in this study. Figure 6.2 also indicates that
fossil energy for biofuel production are highest followed by the fossil energy required for
feedstock production. Transportation energy contributes a small portion to the overall
fossil energy use.
6.3.2 WTP petroleum use
Figure 6.3 shows the WTP petroleum use for the scenarios listed above. In comparing
Figures 6.2 and 6.3, the petroleum contribution to fossil energy use is small indicating that
butanol (and ethanol) can contribute to reducing petroleum use. The WTP petroleum
input requirements for corn butanol and ethanol are comparable due to petroleum not
being used directly in the biofuel production processes
5
but in the activities of corn
farming and transportation which are assumed to be very similar in the life cycles of
4
Notes for WTP Figures: Wu (2008) refers to Wu et al. [22]. The scenarios are: ButOH w/DDGS
= butanol with DDGS co-product credited by the displacement method; ButOH w/DDGS and EtOH
= butanol with DDGS and ethanol credited by the displacement method; ButOH w/DDGS, Acetone
and EtOH = butanol with the three co-products credited by the energy allocation method. The EtOH
results are those for dry mill ethanol in GREET 1.8b
5
Natural gas is the main fossil energy source used in the production of biofuels.
Chapter 6. LCA of Biobutanol for use in Transportation 90
0.0
0.5
1.0
1.5
2.0
2.5
ButOH no
credits
ButOH w/
DDGS
Wu (2008)
ButOH
w/DDGS
and EtOH
ButOH
w/DDGS,
Acetone,
and EtOH
Wu (2008)
w/DDGS,
Acetone,
and EtOH
EtOH no
credits
EtOH w/
DDGS
W
e
l
l

t
o

P
u
m
p

f
o
s
s
i
l

e
n
e
r
g
y

u
s
e
(
M
J
/
M
J

o
f

f
u
e
l

p
r
o
d
u
c
e
d
)
Transport
Production
Farming
WTP with credit
Figure 6.2: WTP fossil energy use (MJ of fossil fuel input/MJ of fuel produced)
the two fuels. Inclusion of co-product credits reduces the WTP petroleum use of corn
butanol by 50% to 75%.
6.3.3 WTP GHG emissions
Figure 6.4 presents the WTP GHG emissions for the various scenarios. A credit is
given for CO
2
uptake during feedstock growth, but this is applied only when the biofuel
is combusted in the vehicle as per GREET convention [12]. Therefore, these WTP
results do not include a CO
2
uptake credit. Corn butanol WTP results range from
60 to 180 g CO
2
-eq./MJ. The lower results are those which assume co-product credits.
Results for ethanol are comparable to those of butanol when co-products are considered.
It is interesting to note that fossil energy use in the most optimistic case for butanol
(considering all co-product credits) is greater than the fossil energy use for the ethanol
production with co-product credits, yet the GHG emissions are greater for ethanol than
butanol in the same scenarios. This result is because butanol consumes more fossil
energy in the form of less carbon intensive natural gas (i.e., during alcohol production)
while ethanol consumes more carbon intensive petroleum during transportation. Thus,
Chapter 6. LCA of Biobutanol for use in Transportation 91
0.0
0.1
0.1
0.2
0.2
0.3
ButOH no
credits
ButOH w/
DDGS
Wu (2008)
ButOH
w/DDGS
and EtOH
ButOH
w/DDGS,
Acetone,
and EtOH
Wu (2008)
w/DDGS,
Acetone,
and EtOH
EtOH no
credits
EtOH w/
DDGS
W
e
l
l

t
o

P
u
m
p

p
e
t
r
o
l
e
u
m

e
n
e
r
g
y

u
s
e
(
M
J
/
M
J

o
f

f
u
e
l

p
r
o
d
u
c
e
d
)
Transport
Production
Farming
WTP with credit
Figure 6.3: WTP petroleum energy use (MJ of petroleum input/MJ of fuel produced)
butanols ability to be transported by pipelines helps in reducing net GHG emissions.
6.3.4 WTW Fossil Energy Use and GHG Emissions
For the WTW results, only the most optimistic scenarios for butanol (i.e., DDGS, ace-
tone, and ethanol co-products credited by the energy allocation method) and ethanol
(i.e., DDGS co-product credited by the displacement method) are compared to that of
conventional gasoline for E85 and Bu85 blends. Figures 6.5 and 6.6 present the WTW
fossil energy use and GHG emissions per km driven by a passenger vehicle fueled by
conventional gasoline (CG), E85, and Bu85.
The trends for fossil energy use and GHG emissions closely match each other, as is
expected from the strong relation between fossil energy combustion and GHG emission.
E85 and Bu85 can oer reductions in fossil energy use by approximately 37-38% when
compared to CG. The fossil energy use and GHG emissions for E85 and Bu85 are almost
the same. Although not graphed, it is evident that when coproduct credits are not
considered for Bu85, then the fossil energy use and GHG emissions would largely exceed
those of CG.
Chapter 6. LCA of Biobutanol for use in Transportation 92
0
20
40
60
80
100
120
140
160
180
200
ButOH no
credits
ButOH w/
DDGS
Wu (2008)
ButOH
w/DDGS
and EtOH
ButOH
w/DDGS,
Acetone,
and EtOH
Wu (2008)
w/DDGS,
Acetone,
and EtOH
EtOH no
credits
EtOH w/
DDGS
W
e
l
l

t
o

P
u
m
p

G
H
G

e
m
i
s
s
i
o
n
s
(
g

C
O
2

e
q
.
/
M
J

o
f

f
u
e
l

p
r
o
d
u
c
e
d
)
WTP with credit
Transport
Production
Farming
Figure 6.4: WTP GHG emissions (CO
2
-eq/MJ of fuel produced)
6.4 Conclusions
The fossil and petroleum energy input as well as GHG emissions resulting from the
production and use of butanol from corn assuming three co-product allocation scenarios
were modeled. The co-product allocation method was found to have considerable impact
on the resulting environmental metrics. Uncertainty in the modeling was not examined
in this work but is a critical issue that should be examined in future research.
Butanol had a high fossil energy input, more than double the input to ethanol as
reported in GREET 1.8b if no co-product allocation was included. The much higher fossil
energy was attributed to the lower yields and higher product recovery energy requirements
for butanol production (i.e., primarily distillation energy). When energy was allocated
to the co-products, results were more in line with those of ethanol. Under all scenarios,
the petroleum contribution to fossil energy use was small, indicating that butanol can
contribute to reducing petroleum use in the transportation sector. The butanol scenarios
presented herein generally reported higher energy use and GHG emissions than those of
Wu et al. [22] due to our lower yield assumptions and higher energy requirements for
product recovery. On a WTW basis Bu85 and E85 can oer reductions in fossil energy
Chapter 6. LCA of Biobutanol for use in Transportation 93
0
1
2
3
4
5
C
G
E
8
5
B
u
8
5
W
e
l
l

t
o

W
h
e
e
l

f
o
s
s
i
l

e
n
e
r
g
y

u
s
e
(
M
J
/
k
m
)
Figure 6.5: WTW fossil energy use (MJ of fossil energy input/km driven)
use and GHG emissions when compared to conventional gasoline. Additional modeling of
butanol production based on state-of-the-art experimental and pilot-scale data is needed,
as is additional data from vehicle testing to facilitate the development of more detailed
WTW analyses.
6.4.1 Recommendations
The environmental performance of corn-butanol is poor because of low butanol yields and
high energy use for recovery (i.e., distillation). The use of lignocellulosic feedstock for
biobutanol production would greatly improve environmental performance, as has been
shown in bioethanol LCA studies [13, 16]. LCA need to be conducted on biobutanol
derived from lignocellulosic feedstocks in order to determine the fuels environmental
performance. Future LCA studies should also include additional functional units, such
as life cycle water use, emissions of volatile organic compounds, and the eects of land
use change.
From a biofuel production perspective, it would benecial to use entire ABE mixture
as a biofuel rather than expending energy to separate the acetone, butanol, and ethanol
products from each other. ABE biofuel is likely to oer a better environmental perfor-
mance due to the higher yield of total solvents and lower energy required for solvent
Chapter 6. LCA of Biobutanol for use in Transportation 94
0
50
100
150
200
250
300
350
C
G
E
8
5
B
u
8
5
W
e
l
l

t
o

W
h
e
e
l

G
H
G

e
m
i
s
s
i
o
n
s
(
g

C
O
2

e
q
.
/
k
m
)
Figure 6.6: WTW GHG emissions (CO
2
-eq/km driven)
recovery; however, this needs to be critically assessed using the LCA methodology. The
feasibility of ABE biofuel and its combustion performance in vehicles would rst need to
be assessed, as is discussed in the following chapter.
A future research project would determine the net GHG emissions, fossil energy use,
and petroleum use associated with the production of ABE biofuel from corn and lignocel-
lulosic feedstock. An ASPEN model can be created to determine the energy requirements
for the ABE production process, and then the GREET software can be used to calculate
the net environmental impacts associated with feedstock (corn and corn stover) produc-
tion, through to conversion of the feedstock to fuel, and nally distribution of the fuel to
retail refueling stations.
The potential use of ABE biofuel would eliminate the need for sequential distilla-
tion to separate butanol from acetone and ethanol; however, water in the fermentation
broth would still require large amounts of energy to remove via distillation. Alternative
separation processes, such as pervaporation, gas stripping, liquid-liquid extraction, and
adsorption-desorption have been pursued [32]. Of these advanced separation processes,
adsorption-desorption requires the least amount of energy. Therefore, future research
should be directed towards developing a novel adsorbent for ABE removal from the fer-
mentation broth. The ideal adsorbent would have the following characteristics: quick
Chapter 6. LCA of Biobutanol for use in Transportation 95
adsorption kinetics, high capacity, low cost, quick desorption kinetics, and ease of re-
generation. Initially, this research should explore various materials (e.g. silicalite and
activated carbon) as adsorbents for ABE mixtures in water. Particular attention should
be paid to determining the eects of material composition, surface area, and particle size.
Literature Cited
[1] K. Sanderson, US biofuels: A eld in ferment, Nature, vol. 444, pp. 673676, 2006.
[2] A. Ragauskas, The path forward for biofuels and biomaterials, Science, vol. 311,
pp. 484489, 2006.
[3] D. Pimentel and T. W. Patzek, Ethanol production using corn, switchgrass, and
wood; biodiesel production using soybean and sunower, Natural Resource Re-
search, vol. 14, pp. 6576, 2005.
[4] T. Patzek, Thermodynamics of the corn-ethanol biofuel cycle, Critical Reviews in
Plant Sciences, vol. 23, pp. 519567, 2004.
[5] M. S. Graboski, Fossil energy use in the manufacture of corn ethanol, Prepared
for the National Corn Growers Association (2002)., Tech. Rep., 2002.
[6] H. Shapouri, J. A. Dueld, and M. Wang, The energy balance of corn ethanol: An
update, U.S. Department of Agriculture Agricultural Economic Report No. 814,
Tech. Rep., 2002.
[7] M. E. D. de Oliveira, B. E. Vaughan, and E. J. Rykiel, Ethanol as fuel: Energy,
carbon dioxide balances, and ecological footprint, Journal of BioScience, vol. 55,
p. 593, 2005.
[8] S. Kim and B. Dale, Ethanol fuels: E10 or E85 - life cycle perspectives, The
International Journal of Life Cycle Assessment, vol. 11, pp. 117121, 2006.
[9] S. Kim and B. Dale, Environmental aspects of ethanol derived from no-tilled corn
grain: fossil energy consumption and greenhouse gas emissions, Biomass and Bioen-
ergy, vol. 28, pp. 475489, 2005.
96
Literature Cited 97
[10] M. Wu, M. Wang, and H. Huo, Fuel-cycle assessment of selected bioethanol pro-
duction pathways in the united states, Centre for Transportation Research Energy
Systems Division,, Argonne National Laboratory, Tech. Rep., 2006.
[11] A. Farrell, R. Plevin, B. Turner, A. Jones, M. Ohare, and D. Kammen, Ethanol
can contribute to energy and environmental goals, Science, vol. 311, pp. 506508,
2006.
[12] Anonymous. Greenhouse gases, regulated emissions, and energy use in
transportation (greet) computer model. Center for Transportation Re-
search Argonne National Laboratory. Argonne, Illinois. [Online]. Available:
http://www.transportation.anl.gov/software/GREET/publications.html
[13] E. Larson, A review of life-cycle analysis studies on liquid biofuel systems for the
transport sector, Energy for Sustainable Development, vol. 2, pp. 109126, 2006.
[14] J. Hill, E. Nelson, D. Tilman, S. Polasky, and D. Tiany, Environmental, economic,
and energetic costs and benets of biodiesel and ethanol biofuels, Proceedings of
the National Academy of Sciences, vol. 103, pp. 11 20611 210, 2006.
[15] H. MacLean, L. Lave, R. Lanky, and S. Joshi, A life-cycle comparison of alternative
automobile fuels, Journal of the Air & Waste Management Association, vol. 50, pp.
17691779, 2000.
[16] S. Spatari, Y. Zhang, and H. MacLean, Life cycle assessment of switchgrass- and
corn stover-derived ethanol-fueled automobiles, Environmental Science and Tech-
nology, vol. 39, pp. 97509758, 2005.
[17] L. P. Koh and J. Ghazoul, Biofuels, biodiversity, and people: Understanding the
conicts and nding opportunities, Biological Conservation, vol. 141, no. 10, pp.
24502460, 2008.
[18] J. Goldemberg and P. Guardabassi, Are biofuels a feasible option? Energy Policy,
vol. 37, pp. 1014, 2009.
[19] T. Searchinger, R. Heimlich, R. A. Houghton, F. Dong, A. Elobeid, J. Fabiosa,
S. Tokgoz, D. Hayes, and T.-H. Yu, Use of US croplands for biofuels increases
greenhouse gases through emissions from land-use change, Science, vol. 319, no.
5867, pp. 12381240, February 2008.
Literature Cited 98
[20] J. Fargione, J. Hill, D. Tilman, S. Polasky, and P. Hawthorne, Land clearing and
the biofuel carbon debt, Science, vol. 319, no. 5867, pp. 12351238, February 2008.
[21] Anonymous, The addition of bio-butanol to GHGenius and a review of the GHG
emissions from diesel engines with urea SCR, S&T Squared Consultants Inc., Tech.
Rep., 2007.
[22] M. Wu, M. Wang, J. Liu, and H. Huo, Assessment of potential life-cycle energy and
greenhouse gas emission eects from using corn-based butanol as a transportation
fuel, Biotechnology Progress, vol. 24, pp. 12041214, 2008.
[23] T. Ezeji, Q. N., and H. Blaschek, Acetone butanol ethanol ABE production from
concentrated substrate: reduction in substrate inhibition by fedbatch technique
and product inhibition by gas stripping, Applied Microbiology and Biotechnology,
vol. 63, pp. 653658, 2004.
[24] T. Ezeji, P. Karcher, N. Qureshi, and H. Blaschek, Improving performance of a
gas-stripping-based recovery system to remove butanol from clostridium beijerinckii
fermentation, Bioprocess and Biosystems Engineering, vol. 27, pp. 207214, 2005.
[25] Environmental management - life cycle assessment - principles and framework, In-
ternational Organization for Standardization (ISO) Std. ISO 14 040, 2006.
[26] Anonymous, Climate change 2007: The physical science basis: Working group
i contribution to the fourth assessment report of the IPCC, IPCC, Cambridge
University Press, Cambridge, Massachusetts, Tech. Rep., 2007.
[27] J. R. Kwiatkowski, A. J. McAloon, F. Taylor, and D. B. Johnston, Modeling the
process and costs of the production of fuel ethanol by the corn dry-grind process,
Industrial Crops and Products, vol. 23, pp. 288296, 2006.
[28] A. McAloon, F. Taylor, and Y. Winnie, A model of the production of ethanol by the
dry grind process, in Proceedings of the Corn Utilization & Technology Conference,
Indianapolis, Indiana, June 2004.
[29] N. Qureshi and H. Blaschek, Economics of butanol fermentation using hyper-
butanol producing clostridium beijerinckii ba101, The Institution of Chemical En-
gineers, vol. 78, pp. 139144, 2000.
Literature Cited 99
[30] T. Ezeji, N. Qureshi, and H. Blaschek, Bioproduction of butanol from biomass:
from genes to bioreactors. Current Opinion in Biotechnology, vol. 18, pp. 220227,
2007.
[31] J. Liu, L. Fan, P. Seib, F. Friedler, and B. Bertok, Downstream process synthesis
for biochemical production of butanol, ethanol, and acetone from grains: Gener-
ation of optimal and near-optimal owsheets with conventional operating units,
Biotechnology Progress, vol. 20, no. 5, pp. 15181527, September-October 2004.
[32] N. Qureshi, S. Hughes, I. S. Maddox, and M. A. Cotta, Energy-ecient recovery
of butanol from model solutions and fermentation broth by adsorption, Bioprocess
and Biosystems Engineering, vol. 27, pp. 215222, 2005.
[33] C. Hernandez-Gaona, J. Cardenas, J. Segovia-Hernandez, S. Hernandez, and
V. Rico-Ramirez, Second law analysis of conventional and nonconventional distil-
lation sequences, Chemical and Biochemical Engineering Quarterly, vol. 19, no. 3,
pp. 235241, September 2005.
[34] M. Collura and W. Luyben, Energy-saving distillation designs in ethanol-
production, Industrial & Engineering Chemistry Research, vol. 27, no. 9, pp. 1686
1696, September 1988.
[35] J. Liu, M. Wu, and M. Wang, Simulation of the process for producing butanol from
corn fermentation, Industrial & Engineering Chemistry Research, vol. 48, no. 11,
pp. 55515557, June 2009.
[36] Anonymous, Model year 2006: fuel economy guide, U.S. Department of Energy
Oce of Energy Eciency and Renewable Energy, Tech. Rep., 2006. [Online].
Available: http://www.fueleconomy.gov
[37] M. Gautam, D. Martin, and D. Carder, Emissions characteristics of higher alco-
hol/gasoline blends, Proceedings of the Institution for Mechanical Engineers, vol.
214, pp. 165182, 2000.
Chapter 7
An Experimental and Kinetic
Modeling Study of Butanol
Combustion
7.1 Introduction
7.1.1 Engine Studies
There have been several engine studies using n-butanol as a fuel or as a blending agent
with gasoline [1, 2, 3, 4, 5, 6]. In the most notable, Yacoub et al. [6] used gasoline
blended with n-butanol) to fuel a single-cylinder SI engine. They found that the butanol
blends had less knock resistance than neat gasoline. The butanol blends also had reduced
CO and hydrocarbon emissions but increased NO
a
emissions. This may be due to the n-
butanol blends having a higher ame temperature and earlier spark timing. Of particular
interest to the present study is that the primary oxygenated hydrocarbon emissions were
butanol, formaldehyde and to a lesser extent, acetaldehyde. A study by Miller et al.
successfully operated unmodied gasoline and diesel engines on blends containing 0-20%
butanol in gasoline and 0-40% n-butanol in diesel fuel [7].
A recent study by Wallner et al. [8] studied 10%butanol-gasoline and 10%ethanol-
gasoline blends in a modern direct-injection four-cylinder SI engine at varying engine
speeds. The dierence in heat release rate was found to be neglible between the biofuel
blends and neat gasoline. At high engine loads, the butanol blend had a reduced knock
resistance, in agreement with the aforementioned study [6]. Due to dierences in energy
100
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 101
density, the brake specic volumetric consumption was the lowest for gasoline, second
lowest for the butanol blend, and highest for the ethanol blend. The regulated emissions
proles for the ethanol and butanol blends were comparable, and the study concluded
that butanol use in existing engines is feasible.
7.1.2 Combustion Chemistry Studies
Predictive models provide a better understanding of the combustion performance and
emissions characteristics of biofuel compositions and why they dier from petroleum
derived materials. The development of a butanol model requires understanding of its
fundamental pyrolysis and oxidation kinetics. However, only a few studies have examined
the combustion chemistry of butanol.
The pyrolysis of n-butanol was studied by Barnard using a static reactor [9]. The
author suggests that pyrolysis is initiated by ssion at the C
3
H
7
-CH
2
OH bond to pro-
duce the n-propyl radical and hydroxymethyl radical. The hydroxymethyl radical further
decomposes to formaldehyde and a hydrogen radical, while the n-propyl radical decom-
poses to ethylene and a methyl radical. In another study, Roberts measured the burning
velocities of n-butanol using shadowgraph images of the ame cone [10]. The results
indicate that the maximum burning velocity of n-butanol is similar to that of n-propanol
and isopentyl alcohol (i.e., approx. 46 cm/sec). A recent study by McEnally and Pfef-
ferle measured temperature and species in atmospheric-pressure coowing laminar non-
premixed methane ames doped with one of four isomers of butanol (including n-butanol)
[11]. They concluded that unimolecular dissociation dominated over H-atom abstraction.
This consisted of C-C ssion followed by /ctc-scission of the resulting radicals. Complex
ssion involving four-center elimination of water was estimated to account for only 1% of
n-butanol decomposition. The most important measured species included propene and
ethylene. Yang and coworkers [12] studied laminar premixed ames fuelled by one of
four isomers of butanol (including n-butanol). Their results identied the combustion
intermediates in the butanol ames, but did not provide concentration proles. The
qualitative data provided lends support to the aforementioned dissociation mechanism
[11].
Recently, Moss et al. published a chemical kinetic mechanism for four butanol iso-
mers (i.e., n-butanol, sec-butanol, iso-butanol, and tert-butanol) and validated it against
ignition delay times measured in a shock tube [13]. Under the given experimental con-
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 102
ditions, the study concluded that n-butanol is mainly consumed by H-atom abstractions
leading to the formation of C
3
CH
7
CHO, acetaldehyde, and ethylene, and, to a lesser
extent, propene and 1-butene. It should be noted that the chemical kinetic mechanism
was not validated against species proles in the shock tube.
Recent work by Dagaut et al. studied the oxidation of 85%butanol-15%gasoline surro-
gate mixtures in a jet stirred reactor (JSR) at 10 atm, equivalence ratios spanning 0.3-2.0,
and temperatures ranging from 770-1270 K [14]. The surrogate mixture was comprised
of iso-octane, toluene, 1-hexene and n-butanol. A novel chemical kinetic mechanism was
derived using mechanisms for each pure component in the butanol-gasoline surrogate
mixture, and it was shown to provide good agreement with the experimental data. The
mechanism indicates that H-abstraction reactions are the main consumption pathway for
the butanol-gasoline surrogate fuel.
The goal of this study is to provide additional experimental data on pure n-butanol
oxidation in several well-dened environments. This study is the rst to model detailed
chemistry in an n-butanol ame. Species and temperature proles are provided for an
n-butanol opposed-ow diusion ame. The paper also presents species proles for an
n-butane opposed-ow diusion ame to elucidate the dierences in combustion between
an alkane and an alcohol. In addition, species proles are presented for n-butanol in a
JSR at 101.325 kPA and 1013 kPa (i.e., 1 atm and 10 atm) with a range of equivalence
ratios (o=0.25-2.0) and temperatures (T)
1
. Burning velocity data is presented for an n-
butanol premixed laminar ame at various equivalence ratios
2
. This comprehensive data
set is used to validate an improved chemical kinetic model of n-butanol oxidation.
7.2 Experimental Methods
7.2.1 Opposed-ow Diusion Flame
A detailed explanation of the experimental opposed-ow diusion ame and correspond-
ing sampling setup was presented in Chapter 4. Briey, the setup consists of two identical
at ame burners with circular burner ports of diameter 25.4 mm, facing each other and
spaced 20 mm apart. A fuel mixture of 94.11% N
2
and 5.89% fuel (99% pure n-butanol
or 99% pure n-butane) was fed through the bottom port at a mass ux rate of 0.0131
1
The JSR experimental data was obtained by M.J. Thomson, P. Dagaut, and C. Togbe
2
The laminar ame speed data was obtained by F. Halter and C. Mouna

m-Rousselle
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 103
g/cm
2
-sec, while an oxidizer mixture of 42.25% O
2
and 57.75% N
2
was fed through the
top port at a mass ux rate of 0.0126 g/cm
2
-sec. At these plug ow conditions, the
Reynolds Number is in the laminar ow regime (i.e. 1c < 400), the ame is on the fuel
side of the stagnation plane, and the fuel side strain rate is approximately 33 :
1
. An
ultrasonic atomizer was used to spray the liquid fuel into a stream of N
2
gas preheated
to 356 K. The gaseous fuel mixture was delivered to the burner via heated stainless steel
tubing. The temperatures of the gases exiting the top and bottom burner ports were
423 K and 356K , respectively. The gas sampling system in these experiments consists
of a quartz microprobe (250 m ID, 300 m OD) connected to a dual-stage pump with
heated heads (388 K) containing PTFE diaphragms. The suction side of the sampling
system consisted of 1,4

tubing and a vacuum pressure gauge connected to the quartz


microprobe. An absolute pressure of 4-6 kPa was measured downstream of the micro-
probe. This was sucient to quench most reactions and ensure accurate data on ame
composition. The compression side of the pump delivered the samples to the analytical
instruments via 1,4

stainless steel tubing heated to 388 K.


Analytical techniques used to analyze the species in the sample included: NDIR for
CO and CO
2
; GC/FID with an HP-Al/S PLOT column for C
1
to C
5
hydrocarbons; and
GC/FID equipped with a methanizer (i.e., Ni catalyst) and Poraplot-U column for bu-
tanol, butanal, acetaldehyde, and formaldehyde. The precision of species measurements
is estimated to be 15%. Temperature measurements were obtained using a 254 jm
diameter wire R-type thermocouple (Pt-Pt/13% Rh) in an apparatus similar to that used
by McEnally et al. [15]. The measured temperatures were corrected for radiation losses.
7.2.2 Jet Stirred Reactor
The JSR experimental setup
3
. used in this study has been described earlier [16, 17]. The
reactor consists of a 4 cm diameter fused silica sphere equipped with four nozzles of 1
mm ID that promote rapid mixing of the gases as they enter the reactor. Prior to the
injectors, the reactants were diluted with high-purity nitrogen (<100 ppm H
2
O, <50 ppm
O
2
, <1000 ppm Ar, <5 ppm H
2
) and mixed. A high degree of dilution (0.1% mol. of fuel)
was used, reducing temperature gradients and heat release in the JSR. The reactants were
high-purity oxygen (99.995% pure) and high-purity n-butanol (99% pure butanol from
3
The JSR setup is located at CNRS, Orleans, France. The experiments were performed by M.J.
Thomson, C. Togbe, and P. Dagaut.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 104
Aldrich), which was sonically degassed before use. The reactants were preheated before
injection to minimize temperature gradients inside the reactor. A Shimadzu LC10 AD
VP pump with an on-line degasser (Shimadzu DGU-20 A3) was used to deliver the fuel to
an atomizer-vaporizer assembly maintained at 473 K. Good thermal homogeneity along
the vertical axis of the reactor (gradients of approximately 1 K/cm) was observed for each
experiment by thermocouple (0.1 mm Pt-Pt/Rh (10%) located inside a thin-wall silica
tube) measurements. The reacting mixtures were sampled by a fused silica low pressure
sonic probe, and then analyzed online by fourier transform infrared spectroscopy (FTIR)
and o-line after collection and storage in 1 L Pyrex bulbs. O-line analyses were done
using gas chromatographs equipped with capillary columns (DB-624 and Carboplot-P7),
a thermal conductivity detector (TCD), and a FID.
The experiments were performed at steady state under constant mean residence times
(t) of 0.07 seconds and 0.7 seconds corresponding to constant pressures (P) of 101.325
kPa (1 atm) and 1013 kPa (10 atm), respectively. The reactants were continually owing
in the reactor, whereas the temperature of the gases inside the JSR was varied stepwise. A
good repeatability was observed in the experiments and reasonable good carbon balance
of 100 15% was achieved.
7.2.3 Laminar Flame Speed Setup
The device to measure the laminar ame speed consists of a stainless steel cylindrical
combustion chamber with an inside volume of 2.4 L
4
. Two tungsten electrodes linked to a
conventional capacitive discharge ignition system, are used to form the spark gap (2.8mm)
at the center of this chamber. Four windows provide optical accesses into the chamber.
The air-fuel mixture was prepared directly in the chamber by adding the fuel and the
air at appropriate partial pressures to reach the total initial pressure. The pressure and
temperature conditions (i.e., 350 K and 90 kPa) were selected to optimize the saturated
vapour pressure of butanol. The chamber was warmed at the desired temperature and an
electric fan, located inside the chamber, mixed all the gases. Gas phase chromatography
analysis was performed to ensure adequate mixing. High-purity n-butanol was injected
by a gasoline injector. The spray of butanol was rapidly vaporised due to the relative
high temperature and the forced convection created by the fan. A delay before ignition
4
The laminar ame speed setup is located at the University of Orleans, France. The experiments
were performed by F. Halter and C. Mouna

m-Rousselle.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 105
avoided any perturbation during the ame propagation. Measurements were limited to
ames having diameters less than 60 mm, implying that the total volume of burned gases
was less than 0.5%. In this initial part of the ame expansion, the total chamber pressure
can be considered constant. Observation times were lower than 15 ms for all the cases
investigated (depending on the equivalence ratio). A range of equivalence ratios, from
0.8 to 1.2, was tested. The error in the ame speed measurements is 2 cm/s.
The image of the ame was obtained by the classical shadowgraphy technique. The
parallel light from an Ar-Ion laser source was created by two plano-convex lenses (25 mm
and 1000 mm focal lengths respectively). The shadowgraphic images were recorded by
using a high speed video CMOS camera
5
operating at 6000 frames per second with an
exposure time of 20 js. The temporal evolution of the expanding spherical ame was
then analyzed to determine the laminar ame speed. A global schematic view of the
system is presented in Figure 7.1.
Figure 7.1: Schematic of the laminar ame speed measurement setup
7.3 Computational Methods
The kinetic modeling for n-butanol oxidation in the three experimental setups was per-
formed using the CHEMKIN modeling package [18]. The JSR was modeled using the
perfectly stirred reactor (PSR) code, the opposed-ow diusion ame was modeled using
the OPPDIF code, and the laminar ame speed was modeled using the premix ame code
5
Photron APX
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 106
(PREMIX). The inputs to each simulation include a detailed chemical kinetic reaction
mechanism, a dataset of thermochemical properties, and a dataset of transport proper-
ties. These input les are available as supplemental material to the journal publication
of this study [19].
7.3.1 Chemical Kinetic Mechanism
The chemical kinetic reaction mechanism developed here is based on a previously pro-
posed mechanism for 85%butanol-15%gasoline surrogate mixtures in a JSR [14]. The
combustion of n-butanol proceeds via unimolecular initiation and hydrogen abstraction
reactions. The fuel radical species formed are consumed via unimolecular decomposition
(/ctc-scission) and bimolecular reactions. Isomerization of radical species is also included
in the mechanism. Table 7.1 presents the structure of species produced during the oxi-
dation of n-butanol. Modications have been made to the original mechanism to provide
better agreement with JSR data at 1 atm and 10 atm and opposed-ow diusion ame
data. Below is a description of the proposed mechanism.
The Dagaut et al. n-butanol oxidation submechanism [14] was built upon a previ-
ously proposed C
1
-C
4
hydrocarbon mechanism [20, 21, 22]. From the C
1
-C
4
mechanism,
Dagaut et al. added 136 reactions to represent the oxidation of n-butanol and the vari-
ous species formed during its decomposition. Due to the absence of fundamental kinetic
studies on n-butanol unimolecular decomposition reactions and abstraction reactions, the
authors allocated reaction rate constants based on published rate data for structurally
similar hydrocarbons and oxygenates. The reaction rates were allocated to provide better
agreement with JSR data for 85%butanol-15%gasoline surrogate mixtures at 10 atm [14].
The present author (i.e., S.M. Sarathy) has modied Dagaut et al.s mechanism to
better predict the experimentally measured opposed-ow diusion ame species proles.
The revised mechanism consists of 878 reactions involving 118 species. The following is
a description of specic changes made to the Dagaut et al. mechanism.
The simulations using the OPPDIF code were unable to converge upon a solution
using the aforementioned C
1
-C
4
hydrocarbon mechanism. In order to obtain convergence,
the following modications were made:
removal of the reaction C
3
H
6
+

C
2
H

BUTYNE +

CH,
replacement of the reaction n

C
3
H
7
+ (M)

C
3
H
6
+

H + (M) by the reaction
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 107
Table 7.1: Chemical structures of species during the oxidation of n-butanol
n

C
3
H
7

C
3
H
6
+

H with reaction rate constants proposed by Curran [23].
This study has also made the following modications to the former n-butanol sub-
mechanism, in order to provide better agreement with the experimental data presented
herein:
The reaction rate constant
6
for the reaction C
4
H
9

O


H + C
3
H
7
CHO has been
changed to
8.89r10
10
1
0.75
exp
(
21060
co|
nc|
11
)
c:
3
:o| :
based on the rate expression recommended by Curran for C
3
H
7

O


H + C
2
H
5
CHO
[23].
6
Units are presented in cal, K, mol, cm
3
, and s according to the CHEMKIN convention
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 108
The reaction rate constant for the reaction C
4
H
9
OH +

O


OH + C
4
H
9

O has
been changed to
1.58r10
10
1
2.00
exp
(
448
co|
nc|
11
)
c:
3
:o| :
based on the rate expression recommended by Marinov for the ethanol reaction
C
2
H
5
OH +

O


OH +C
2
H
5

O [24].
The reaction rate constant for the reaction C
4
H
9
OH +

H

H
2
+ C
4
H
9

O has been
changed to
5.36r10
4
1
2.53
exp
(
8754
co|
nc|
11
)
c:
3
:o| :
based on the rate expression recommended by Park et al. for the ethanol reaction
C
2
H
5
OH +

H

H
2
+ C
2
H
5

O [25].
The reaction rate constant for the reaction C
4
H
9
OH +

OH

H
2
O + C
4
H
9

O has
been changed to
7.46r10
11
1
0.30
exp
(
1634
co|
nc|
11
)
c:
3
:o| :
based on the rate expression recommended by Park et al. for the ethanol reaction
C
2
H
5
OH +

OH

H
2
O + C
2
H
5

O [25].
The reaction rate constant for the reaction C
4
H
9
OH +

H

H
2
+ a

C
4
H
8
OH has
been changed to
2.58r10
7
1
1.65
exp
(
2827
co|
nc|
11
)
c:
3
:o| :
based on the rate expression recommended by Marinov for the ethanol reaction
C
2
H
5
OH +

H


OH +

C
2
H
4
OH [24].
The reaction rate constant for the reaction C
4
H
9
OH +

OH

H
2
O + a

C
4
H
8
OH
has been changed to
4.64r10
11
1
0.15
c:
3
:o| :
based on the rate expression recommended by Marinov for the ethanol reaction
C
2
H
5
OH +

OH

H
2
O +

C
2
H
4
OH [24].
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 109
The original mechanism [14] employed the same reaction rate constants for uni-
molecular dissociation at the C
2
H
5
-C
2
H
4
OH and C
3
H
7
-CH
2
OH bond sites. How-
ever, the reaction rate constant for the reaction C
4
H
9
OH


C
2
H
5
+

C
2
H
4
OH has
been changed to
5.0r10
16
1 exp
(
86221
co|
nc|
11
)
c:
3
:o| :
(7.1)
since the BDE of the C
2
H
5
-C
2
H
4
OH bond is lower than the BDE of the C
3
H
7
-
CH
2
OH due to the proximity of the OH group in the latter.
7.3.2 Thermochemical Data
The original thermochemical data for the butanol related species was calculated using
the software THERGAS [26], based on the group and bond additivity methods proposed
by Benson [27].
7.3.3 Transport Properties
In certain combustion applications, such as the JSR, the overall rate is assumed to
be kinetically controlled since the transport processes occur innitely fast. Therefore,
the original Dagaut et al. mechanism [14] did not include transport properties of any
species. However, the transport processes are rate-controlling in laminar diusion ames.
This study obtained the molecular transport parameters for species using a variety of
methods. The transport properties for the majority of compounds were already available
in the previously published C
1
-C
4
hydrocarbon mechanism [20, 21, 22]. In addition, the
transport properties of several species were obtained from a study by Gail et al. on
the combustion of methyl butanoate [28]. The transport properties of species with no
previously published data were determined as follows. For stable species, this study used
the correlations developed by Tee, Gotoh, and Stewart [29], as described in Wang and
Frenklach [30], to calculate the Lennard-Jones collision diameter and potential well depth
using the critical pressure (P
c
), critical temperature (T
c
), and boiling point (T
o
) of the
species. The P
c
, T
c
, and T
o
for stable species were obtained from the NIST Chemistry
WebBook [31]. The polarizability in cubic Angstroms of stable species was obtained from
the CRC Handbook of Chemistry and Physics [32]. The dipole moment was obtained
from [33]. The index factor which describes the geometry of the molecule was determined
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 110
from the molecular structure. For new radical species, the aforementioned literature data
is not readily available, so the transport properties of their stable counterpart were used.
7.4 Results and Discussion
7.4.1 Jet Stirred Reactor
The JSR
7
allows studying n-butanol oxidation in a ameless premixed environment. The
concentration of species at each equivalence ratio and temperature condition in the JSR
was measured by sonic probe sampling and GC and FTIR analyses. The measured
species included hydrogen (H
2
), oxygen (O
2
), water (H
2
O), carbon monoxide (CO), car-
bon dioxide (CO
2
), methane (CH
4
), acetylene (C
2
H
2
), ethylene (C
2
H
4
), ethane (C
2
H
6
),
propene (C
3
H
6
), acetaldehyde (CH
3
CHO), formaldehyde (CH
2
O), butanal (C
3
CH
7
CHO),
1-butene (1-C
4
H
8
), and n-butanol (C
4
H
9
OH).
The following oxygenated products were detected: ethyloxirane (C
4
H
8
O), propanal
(C
3
CH
5
CHO), 2-propenal (C
2
CH
3
CHO), methyloxirane (C
3
H
6
O), oxirane (C
2
H
4
O), bu-
tanal, formaldehyde and acetaldehyde. The oxiranes, 2-propenal, and propanal are
formed at ppm levels, and therefore no concentration proles are reported. Enols, which
are unsaturated alcohols, were not detected. A comparison with results obtained for
ethanol in similar conditions and keeping the initial carbon content constant shows bu-
tanol oxidation produces less aldehydes overall. The maximum amount of acetaldehyde
production is reduced by approximately 70% when changing the fuel from ethanol to
butanol.
Initially, the chemical kinetic mechanism was used to predict JSR data at 1013 kPa
and three equivalence ratios (i.e., o=0.5, o=1.0, and o=2.0). Figures displaying the
modeling (open symbols with line) and experimental results (solid symbols) for all the
experimental conditions are not presented here, but are available in Appendix B. Fol-
lowing is a comparison of the experimental data and the proposed model for the o=1.0
equivalence ratio. Unless otherwise mentioned, similar trends were observed at other
equivalence ratios.
The gures are plotted on a logarithmic y-axis and the models ability to reproduce the
experimental data is discussed qualitatively and quantitatively. The models prediction is
7
The JSR setup is located at CNRS, Orleans, France. The experiments were performed by M.J.
Thomson, C. Togbe, and P. Dagaut.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 111
considered good if the shape of the model prole closely matches the experimental prole,
and if the predicted maximum mole fraction is within a factor 2 of the measured maximum
mole fraction. Figure 7.2 presents the experimental measurements (solid symbols) and
modeling results (open symbols with line) of n-butanol obtained at P=1013 kPa and
o=1.0. The experimental results show that with increasing temperature, the n-butanol
levels drop signicantly between 800 K and 900 K. This corresponds to a large increase
in the concentrations of C
3
CH
7
CHO, 1-C
4
H
8
, and C
3
H
6
, all of which are products of H
abstraction pathways. The concentrations of these compounds then quickly decrease as
the temperature increases. C
2
H
4
, C
2
H
6
, CH
3
CHO, and CH
2
O concentrations are also
shown to increase between 800 K and 900 K. However, as the temperature increases
further, the concentrations of these species tends to diminish at a slower rate than the
aforementioned species.
The model predictions (open symbols with line) for P=1013 kPa, o=1.0 are also
shown in Figure 7.2. Reasonably good agreement is obtained for all measured species.
The major product species (i.e., CO, CO
2
, and H
2
O) are well predicted by the model.
CH
4
, C
2
H
6
, C
2
H
4
, H
2
, and CH
2
O are also reasonably well predicted across the entire
temperature range. The reactivity of n-butanol is well predicted between 800 K and 950
K, but at greater temperatures the reactivity is overpredicted. Species concentrations of
C
3
CH
7
CHO, C
3
H
6
, 1-C
4
H
8
, and CH
3
CHO are well predicted until approximately 1000 K,
above which they become underpredicted. C
2
H
2
concentrations are only well predicted
at intermediate temperatures (i.e., 950-1050 K).
To further validate the model, jet stirred reactor data set for n-butanol oxidation is
presented at 101.3 kPa, four equivalence ratios (i.e., o=0.25, o=0.5, o=1.0, and o=2.0)
and a range of temperatures between 800-1250 K. Temperatures below 800 K are not
presented since the fuel was not found to react at these lower temperatures.
Figure 7.3 presents the experimental measurements and modeling results of n-butanol
obtained at o=1.0. The experimental results (solid symbols) show that with increasing
temperature, the n-butanol levels drop signicantly between 950 K and 1050 K. This
corresponds to an increase in the concentrations of C
3
CH
7
CHO and propene, which are
products of H abstraction pathways. The C
3
CH
7
CHO concentration peaks around 980 K
and then decreases quickly, while propene concentration peaks at approximately 1060 K.
The concentrations of CO, CO
2
, CH
4
, C
2
H
4
, C
2
H
6
, C
2
H
2
, CH
3
CHO, H
2
O, H
2
, 1-C
4
H
8
,
and CH
2
O concentrations are also shown to increase above 950 K. The model predictions
(open symbols with line) for o=1.0 are also shown in Figure 7.3. The reactivity of
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 112

1E-06
1E-05
1E-04
1E-03
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C4H8
C4H9OH
C3H7CHO
C3H6

1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
CO
CO2
CH4

1E-06
1E-05
1E-04
1E-03
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C2H2
C2H4
CH3CHO
C2H6

1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
H2O
H2
CH2O

Figure 7.2: Comparison of the experimental and predicted concentration proles obtained
from the oxidation of n-butanol in a JSR at o=1, P=1013 kPa, t=0.7 s.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 113
n-butanol is well predicted below 950 K, but at higher temperatures the reactivity is
underpredicted by the model. Due to this lower reactivity, the models prediction of
C
3
CH
7
CHO is underpredicted and shifted towards higher temperatures. H
2
O is well
predicted by the model at all temperatures, while CO
2
and CH
2
O are underpredicted
below 1150 K. CH
4
, C
2
H
6
, 1-C
4
H
8
, and O
2
are also reasonably well predicted across the
entire temperature range. CH
3
CHO and C
3
H
6
are well predicted below 1100 K, but are
overpredicted at higher temperatures. The model underpredicts the concentration CO,
H
2
, C
2
H
4
, and C
2
H
2
across most of the temperature range.


1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C4H9OH
C3H7CHO
C3H6
1-C4H8

1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

CO
CO2
CH4

1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6

1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

H2O
H2
CH2O
O2

Figure 7.3: Comparison of the experimental and predicted concentration proles obtained
from the oxidation of n-butanol in a JSR at o=1, P=101.3 kPa, t=0.07 s.
Figure 7.4 presents the experimental and modeling results obtained at o=2.0 in the
JSR. The experimental data shows a similar trend as observed at o=1.0. The models pre-
diction of n-butanol reactivity is underpredicted above 1000 K. In addition, the C
3
H
7
CHO
concentrations are poorly predicted by the model across the entire temperature range.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 114
The concentrations of CO, CH
4
, H
2
O, C
2
H
6
, C
2
H
4
, C
2
H
2
, 1-C
4
H
8
, and O
2
are well pre-
dicted by the model. CO
2
, CH
2
O, and H
2
are generally underpredicted by the model
across the range of temperatures studied, while CH
3
CHO and C
3
H
6
are well predicted
below 1100 K and overpredicted at higher temperatures.

1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C4H9OH
C3H7CHO
C3H6
1-C4H8

1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

CO
CO2
CH4

1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6

1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

H2O
H2
CH2O
O2

Figure 7.4: Comparison of the experimental and predicted concentration proles obtained
from the oxidation of n-butanol in a JSR at o=2, P=101.3 kPa, t=0.07 s.
Figures 7.5 and 7.6 present the modeling and experimental data obtained at o=0.5
and o=0.25, respectively. Under these fuel lean conditions, the experimental data shows
similar trends to the higher equivalence ratios, except considerably lower concentrations
of C
2
H
2
are detected. Model predictions for C
2
H
2
are below 10 ppm, and therefore
not included in the gures. The model still underpredicts the reactivity of n-butanol
above 1000 K, and its prediction of the C
3
H
7
CHO prole is shifted towards higher tem-
peratures. The model performs well at predicting the concentrations of CH
2
O, H
2
O,
CH
3
CHO, C
3
H
6
, C
2
H
6
, CH
4
, CH
2
O, 1-C
4
H
8
, O
2
, and C
2
H
4
. As observed at o=1.0, the
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 115
concentration of CO
2
is underpredicted below 1150 K. CO is well predicted below 1000 K,
but the maximum mole fraction is underpredicted. H
2
concentrations are well predicted
at o=0.25, but at o=0.5 the maximum mole fraction is underpredicted.

1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C4H9OH
C3H7CHO
C3H6
1-C4H8

1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

CO
CO2
CH4

1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6

1E-05
1E-04
1E-03
1E-02
1E-01
1E+00
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

H2O
H2
CH2O
O2

Figure 7.5: Comparison of the experimental and predicted concentration proles obtained
from the oxidation of n-butanol in a JSR at o=0.5, P=101.3 kPa, t=0.07 s.
Sensitivity analyses and reaction path analyses were conducted to interpret the mod-
eling results at 101.3 kPa. Reaction paths analyses were performed for o=1.0, o=2.0,
o=0.5, and T=1160 K using the normalized reaction rates (refer to Figure 7.7). Accord-
ing to the proposed model at o=1.0, the leading consumption pathway is complex ssion
leading to the formation of 1-C
4
H
8
and H
2
O (25%). H-atom abstraction is responsible
for consuming more than 60% of the n-butanol, with H atoms (29%) and OH (57%)
radicals being the main contributors. At the given temperature, over 35% of the initial
n-butanol goes to form CH
3
CHO, C
2
H
5
radical, C
3
H
6
, and CH
2
OH radical, while over
40% leads to the formation of 1-C
4
H
8
. Most of the 1-C
4
H
8
breaks down to form C
3
H
6
via
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 116

1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C4H9OH
C3H7CHO
C3H6
1-C4H8

1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

(
P
P
M
)
CO
CO2
CH4

1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6

1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
H2O
H2
CH2O

Figure 7.6: Comparison of the experimental and predicted concentration proles obtained
from the oxidation of n-butanol in a JSR at o=0.25, P=101.3 kPa, t=0.07 s.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 117
Table 7.2: Comparison of maximum measured and predicted product concentrations
(ppm) in the JSR at P=101.3 kPa, t=0.07s. Italicized numbers represent measured
values, bold numbers represent predicted values, and the shaded column is the ratio of
measured to predicted.
Measured Parameter =0.25 = 0.5 = 1.0 = 2.0
5397 5363 4794 1796
4911 4940 4745 1696 H
2
O
1.1 1.1 1.0 1.1
3126 3833 1970 279
3425 3517 3281 125 CO
2

0.9 1.1 0.6 2.2
2814 1499 2533 2223
1147 730 1137 1652 CO
2.5 2.1 2.2 1.3
223 206 184 163
162 139 123 108 CH
2
O
1.4 1.5 1.5 1.5
147 203 264 353
131 131 272 334 CH
4

1.1 1.6 1.0 1.1
12 23 84 349
7 6 37 194 C
2
H
2

1.6 3.9 2.3 1.8
531 684 750 884
367 346 506 581 C
2
H
4

1.4 2.0 1.5 1.5
27 50 75 90
31 42 74 92 C
2
H
6

0.9 1.2 1.0 1.0
121 116 118 132
144 158 181 207 C
3
H
6

0.8 0.7 0.7 0.6
49 46 46 44
55 49 44 41 1-C
4
H
8

0.9 0.9 1.0 1.1
55 75 84 74
92 88 83 77 CH
3
CHO
0.6 0.8 1.0 1.0
57 47 27 38
37 32 27 23 C
3
H
7
CHO
1.5 1.5 1.0 1.6
492 780 1498 2271
378 325 603 1042 H
2

1.3 2.4 2.5 2.2
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 118
the aC
3
H
5
radical. Figure 7.7 also indicates that unimolecular decomposition reactions
become more important as the equivalence ratio is increased, but H-atom abstraction
reactions are still predominant. As the equivalence ratio is decreased, the unimolecular
decomposition reaction become negligible and the fuel is consumed by H-atom abstrac-
tion reactions.
Besides CO, CO
2
, and H
2
O, the most abundant measured species at o=1.0 were C
2
H
4
,
H
2
, and CH
4
. The proposed model indicates that H
2
is largely formed via the following
pathways:
1. H abstraction from CH
2
O (28%) which is mainly formed by the eventual decom-
position of the cC
4
H
8
OH fuel radical via CH
2
OH;
2. H abstraction from the fuel (22%).
The leading pathway to C
2
H
4
formation is decomposition of the C
2
H
5
radical (46%)
which is formed via decomposition of the fuel radicals, dC
4
H
8
OH and aC
4
H
8
OH. Another
major pathway is via CH
3
abstraction from C
3
H
6
(16%), which is a major product of
1-C
4
H
8
decomposition and cC
4
H
8
OH fuel radical decomposition. The model indicates
that CH
4
is mainly produced via H abstraction by methyl radicals from CH
2
O (30%),
C
2
H
4
(22%), and H
2
(12%).
As mentioned above, the model underpredicts n-butanol reactivity at higher temper-
atures. A sensitivity analysis was conducted at 1160 K to determine which reactions
have a large aect on n-butanol concentration. A positive sensitivity coecient implies
that an increase in the reactions forward rate will increase the n-butanol concentration
at the specied temperature and equivalence ratio. Figure 7.8 displays the sensitivity
coecients in the JSR at the four equivalence ratios studied. There is a strong negative
sensitivity to the chain-branching reaction H+O
2
=O+OH across all the equivalence ra-
tios. At the fuel lean equivalence ratios, chain branching reactions become increasingly
important since H-atom abstraction reactions are dominant. However, the inability of
the model to well predict the n-butanol reactivity is unlikely due to errors in the reac-
tions rates of these extensively studied reactions. The sensitivity analysis indicates that
at o=1.0 and o=2.0, unimolecular decomposition of the fuel is important. In addition,
C
3
H
6
chemistry plays an important role as a key intermediate and nal stable product,
as shown in the rate of production analysis. At o=0.5 and o=0.25, H-atom abstrac-
tion reactions from the fuel show strong negative sensitivities. Therefore, the proposed
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 119



OH
C

OH
C

OH
C

OH
C

OH
O

C

OH
CH
3

OH C

C H
3
CH
2

C H
3
CH
2

O
C

OH
O
H

OH

+
13%
7%
16%
8%
16%
+
25%
7%
5%
+
95%
96%
+
97%
+
aC
4
H
8
OH
bC
4
H
8
OH
dC
4
H
8
OH
cC
4
H
8
OH
C
4
H
9
O
H
2
O
+ 100%
+
87%
8%
31%
6%
14%
7%
12%
6%
13%
0%
0%
0%
26%
18%
26%
13%
17%

Figure 7.7: Reaction pathway diagram for n-butanol oxidation in the JSR at o=0.5
(normal text), o=1 (bold text),o=2 (italicized text), P=101.3 kPa, t=0.07s, and T=1160
K.
model would benet from fundamental rate studies on n-butanol H-atom abstraction and
unimolecular decomposition reactions.
7.4.2 Opposed-ow Diusion Flame
The proposed model was further validated against experimental data obtained in an
opposed-ow diusion ame. The opposed-ow diusion ame allows us to study the
oxidation of n-butanol in a non-premixed ame environment. Concentration proles
for species were obtained by sampling the product gas at various points between the two
burner ports and then analyzing it using a variety of analytical techniques. The measured
species included n-butanol (C
4
H
9
OH), carbon monoxide (CO), carbon dioxide (CO
2
),
methane (CH
4
), acetylene (C
2
H
2
), ethylene (C
2
H
4
), ethane (C
2
H
6
), propane (C
3
H
8
),
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 120





-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4
HCO+M<=>H+CO+M
H+O2<=>OH+O
C4H9OH=>CH3+CC3H6OH
C4H9OH=>C2H5+C2H4OH
C4H9OH+OH=>H2O+BC4H8OH
C4H9OH+H=>H2+C4H9O
C4H9OH+H=>H2+AC4H8OH
C4H9OH(+M)<=>C4H8+H2O(+M)
C3H6+H<=>AC3H5+H2
AC3H5+H(+M)<=>C3H6(+M)
Normalised Sensitivity Coefficient
phi = 0.25
phi = 0.5
phi = 1.0
phi = 2.0

Figure 7.8: Sensitivity of n-butanol concentraion to select reactions in the JSR at o=1.0,
P=101.3 kPa, t=0.07 s, and T=1160 K.
propene (C
3
H
6
), propyne (pC
3
H
4
), acetaldehyde (CH
3
CHO), formaldehyde (CH
2
O), bu-
tanal (C
3
CH
7
CHO), 1-butene (1-C
4
H
8
), 1,3-butadiene (1,3-C
4
H
6
, cis-2-butene ((Z)-2-
C
4
H
8
), and cis-2-butene ((E)-2-C
4
H
8
). Species below the experimental limit of quan-
tication (5 ppm) included 1-butyne (1-C
4
H
6
), n-butane (C
4
H
10
), 1-pentene C
5
H
10
, n-
pentane (C
5
H
12
), n-hexane (C
6
H
14
), 1-hexene C
6
H
12
, and benzene (C
6
H
6
). Measurable
levels of acrolein, acetone, ethanol, methanol, and unsaturated alcohols were not detected
in the ame.
Figure 7.9 displays the experimentally measured (solid symbols) and model predicted
(open symbols with line) species and temperature proles obtained in the opposed-ow
diusion ame
8
. (Z)-2-C
4
H
8
and E)-2-C
4
H
8
not plotted since both maximum measured
and predicted concentrations were below 100 ppm. The experimental results (solid sym-
bols) show that the n-butanol concentration begins decreasing quickly at a distance of 4
mm from the fuel port. As the fuel is consumed, the CO and CO
2
concentrations begin
8
Larger gures are available in Appendix B
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 121
rising. All of the n-butanol is consumed at a distance of approximately 7 mm from the
fuel port, which corresponds closely to the visually observed ame front. Just before
the ame front, at around 6.5 mm from the fuel port, the concentrations of hydrocarbon
species reach their maximum. Besides CO and CO
2
, the most abundant measured species
are C
2
H
4
, C
2
H
2
, CH
4
, and C
3
H
6
. Thus, there is a general agreement with the o=1.0 JSR
data which also showed high levels of C
2
H
4
and CH
4
.
In order to elucidate the dierences in combustion between alkane combustion and
alcohol, opposed-ow diusion ame proles were also generated for n-butane under sim-
ilar conditions. The supplemental material contains ame proles for measured species
and model predictions using the present mechanism with n-butane as the fuel. Ta-
ble 7.3 displays the maximum measured mole fractions in the n-butanol and n-butane
ames. Considering that the experimental error is 15%, the two ames have similar
concentrations of CO
2
, CO, CH
4
, C
2
H
6
, C
2
H
4
, C
2
H
2
, C
3
H
6
, and C
3
H
8
. The n-butanol
ame has higher measured levels of CH
2
O, CH
3
CHO, C
3
H
7
CHO, 1-C
4
H
8
. Modeling
predictions for both the ames yield similar results. The higher levels of oxygenated
compounds are expected in the n-butanol ame since the fuel is oxygenated. 1-C
4
H
8
is
formed during the combustion of n-butanol via important unimolecular decomposition
and via H-atom abstraction pathways, as is shown in the discussion below. n-Butane
reacts via H-abstraction from primary and secondary carbon atoms to form 1-C
4
H
9
and
2-C
4
H
9
radicals, unimolecular decomposition to form C
2
H
5
radicals, and unimolecular
decomposition to form C
3
H
7
and CH
3
radicals. The 1-C
4
H
9
radicals primarily undergo
-scission to form C
2
H
2
and C
2
H
5
, while 2-C
4
H
9
radicals break down to produce C
3
H
6
,
1-C
4
H
8
, (Z)-2-C
4
H
8
, and/or (E)-2-C
4
H
8
. The pathways leading to 1-C
4
H
8
in n-butanol
are more direct than the pathways in n-butane, thus rationalizing the higher measured
and predicted 1-C
4
H
8
levels in the n-butanol ame.
The models prediction (open symbols with line) of temperature species proles in
the opposed-ow diusion is also shown in Figure 7.9. The model well-reproduces the
experimentally measured temperature prole. The reactivity of n-butanol is also well
predicted by the model. The maximum concentration of CO
2
is well-predicted by the
model, while the maximum concentration of CO is underpredicted by approximately
0.6%. There is a shift in the measured CO and CO
2
proles towards from the fuel port.
This shift is caused due to ow eld disturbances introduced by the quartz probe used in
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 122

0
200
400
600
800
1000
1200
1400
1600
1800
2000
0 5 10 15 20
DISTANCE FROM FUEL PORT (mm)
T
E
M
P
E
R
A
T
U
R
E

(
K
)
measured
corrected
predicted

0%
2%
4%
6%
8%
10%
12%
0 2 4 6 8 10 12 14 16 18 20
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
%
)
CO2
CO
C4H9OH

0
2000
4000
6000
8000
10000
12000
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
p
p
m
)
C2H4
C2H2
CH4

0
500
1000
1500
2000
2500
3000
3500
4000
4500
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
p
p
m
)
C2H6
C4H8
C3H6

0
50
100
150
200
250
300
350
400
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
p
p
m
)
C3H7CHO
C3H8
C3H4
1,3-C4H6

0
200
400
600
800
1000
1200
1400
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
p
p
m
) CH2O
CH3CHO

Figure 7.9: Experimental and computed proles obtained from the oxidation of n-butanol
in an atmospheric opposed-ow ame (5.89% C
4
H
9
OH, 42% O
2
).
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 123
Table 7.3: Comparison of maximum measured species in n-butanol (C
4
H
9
OH) and n-
butane (C
4
H
10
) opposed-ow diusion ames.

Measured Parameter C
4
H
9
OH C
4
H
10

CO
2
Carbon Dioxide (%) 10 9.7
CO Carbon Monoxide (%) 4.0 3.9
CH
4
Methane (ppm) 3621 2608
CH
2
O Formaldehyde (ppm) 592 366
C
2
H
6
Ethane (ppm) 1035 1068
C
2
H
4
Ethylene (ppm) 10260 8862
C
2
H
2
Acetylene (ppm) 3964 2864
C
3
H
6
Propylene (ppm) 1750 1420
C
3
H
8
Propane (ppm) 119 75
1-C
4
H
8
1-Butene (ppm) 852 101
CH
3
CHO Acetaldehyde (ppm) 1173 23.9
C
3
H
7
CHO Butyraldehyde (ppm) 49 below detection limit
these experiments. The probe
9
had a large inner diameter and short tip which caused
the ame to be drawn towards the fuel port. This probe eect was resolved by moving
to a new probe design
10
.
The model performs well qualitatively, in that it well reproduces the shape of the
experimental proles; however, there is a shift in the measured proles towards from
the fuel port. In the following discussion, the models prediction is considered good if
predicted maximum mole fraction is within a factor 1.5 of the measured maximum mole
fraction. The model performs well at predicting the maximum concentrations of C
2
H
4
,
C
2
H
6
, 1-C
4
H
8
, and CH
3
CHO. The model moderately overpredicts (1.5-2.5 times) the
concentration of CH
4
, C
2
H
2
, pC
3
H
4
, C
3
H
6
, and CH
2
O. There are large overpredictions
(greater than 5 times) of 1,3-C
4
H
6
and C
3
H
7
CHO, and large underpredictions of C
3
H
8
(6 times). However, these compounds are only found in small quantities in both the
modeling and experimental results.
Another measure of the models qualitative performance is to compare the relative
concentration of species. Although the model poorly predicts the maximum concentra-
9
This probe was made at the Department of Chemistry Glass Blowing Shop
10
The new probe design is described in Chapter 4 and improved results for methyl decanoate opposed-
ow diusion ames are shown in Chapter 9
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 124
tions of several species, the relative concentration of species is reasonably well reproduced
by the model. For example, both the experimental data and model predictions show the
minor hydrocarbon species in order of decreasing maximum concentration are C
2
H
4
,
C
2
H
2
, CH
4
, C
3
H
6
, C
2
H
6
, 1-C
4
H
8
, 1,3-C
4
H
6
, pC
3
H
4
, C
3
H
8
. For oxygenated compounds,
both the experimental data and model predicted values show that maximum mole frac-
tions decrease in the order of CH
3
CHO, CH
2
O, and C
3
H
7
CHO.
Sensitivity analyses and reaction path analyses were performed to interpret the mod-
eling results in the opposed-ow diusion ame. Reaction path analyses were performed
at three dierent temperatures, as follows:
low temperature, T=858 K, corresponding to 5.25mm from the fuel port and 34%
of the fuel consumed;
intermediate temperature, T=1170 K, corresponding to 6.38mm from the fuel port
and 78% of the fuel consumed;
and high temperature, T=1520 K, corresponding to 7.5mm from the fuel port and
99% of the fuel consumed.
Figure 7.10 displays the primary reactions paths involved in the consumption of n-
butanol at the three aforementioned temperatures, with italicized, regular, and bold texts
referring to low, intermediate, and high temperature conditions, respectively. Percentages
corresponding to each pathway are rounded to the nearest whole number.
The low temperature analysis provides an understanding of how fuel consumption
initiates in the ame. According to the proposed model, H-atom abstraction accounts
for the nearly 100% of the fuel consumption, with abstraction being dominated by H
atoms (43%), OH radicals (22%), CH
3
radicals (11%), and aC
3
H
5
radicals (9%). At
intermediate temperatures, the analysis provides a comparison to the JSR study, while
providing an understanding of reactions occurring near the ame front. As in the JSR,
fuel consumption is dominated by H-atom abstraction reactions (70%) by H atoms (55%)
and OH radicals (25%). At high temperatures, H-abstraction reactions are negligible and
the energy is now available to activate unimolecular decomposition reactions. However,
the amount of fuel that reaches these conditions is small, and therefore these reactions
do not contribute signicantly to the formation of product species.
The fuel radicals formed via H-atom abstraction are primarily consumed via -
scission. At the low temperature condition, over 60% of the initial fuel goes to form
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 125
CH
3
CHO, C
2
H
5
radical, C
3
H
6
, and CH
2
OH radical. Another 38% goes to form C
3
H
7
CHO,
1-C
4
H
8
, H atoms, and OH radicals. At the intermediate temperature condition, H-atom
abstraction reactions are still predominant, but complex ssion leading also plays a role
(17%), and 35% of the fuel goes to form 1-C
4
H
8
. As the temperature increases, the
1-C
4
H
8
is consumed to form C
3
H
6
via the aC
3
H
5
radical.
Besides CO, CO
2
, and H
2
O, the most abundant measured species was C
2
H
4
. At
its maximum predicted concentration (T=1280 K, distance=6.75 mm), the proposed
model indicates that C
2
H
4
is formed via decomposition of the C
2
H
5
radical (39%), which
was formed via decomposition of the fuel radicals (53%), dC
4
H
8
OH and aC
4
H
8
OH, and
unimolecular decomposition of the fuel (32%). The three other pathways contributing
to more than 10% of the total ux are via -scission of the cC
3
H
6
OH radical (17%), -
scission of the C
2
H
4
OH radical (13%), and via CH
3
abstraction from C
3
H
6
(12%). Both
the cC
3
H
6
OH radical and the C
2
H
4
OH radical are formed via unimolecular decomposition
of the fuel (refer to Figure 7.10)
A sensitivity analysis was conducted for n-butanol at the low and intermediate tem-
peratures. Figure 7.11 displays the reactions to which the n-butanol concentration is
sensitive to under these conditions. Strong negative sensitivity coecients are observed
for unimolecular decomposition reactions that are responsible for consuming the fuel.
Increasing the forward rates of these reactions will serve to decrease the fuel concen-
tration. Negative sensitivities are also observed for H abstraction from the , o, and
carbons, which were shown to be major pathways for fuel consumption in the reaction
path analysis. As in the JSR, it can be concluded that the proposed model would ben-
et from fundamental rate studies for n-butanol H-atom abstraction and unimolecular
decomposition reactions. Positive sensitivity coecients are observed for several C
2
and
C
3
reactions that compete for H and CH
3
radicals.
7.4.3 Laminar Flame Speed
The proposed model was also validated against experimentally measured laminar burn-
ing velocities
11
. This experiment allows the study of the combustion of butanol in a
premixed ame environment. The laminar burning velocity was determined over a range
of equivalence ratios and was compared to the data previously obtained by Roberts [10].
11
The laminar ame speed setup is located at the University of Orleans, France. The experiments
were performed by F. Halter and C. Mouna

m-Rousselle.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 126

OH
C

OH
C

OH
C

OH
C

OH
O

C

OH
CH
3

OH C

C H
3
CH
2

C H
3
CH
2

O
C

OH
OH

O
H

+
16%
20%
25%
14%
25%
+
0%
0%
0%
+
99%
99%
+
99%
+
aC
4
H
8
OH
bC
4
H
8
OH
dC
4
H
8
OH
cC
4
H
8
OH
C
4
H
9
O
+
95%
+
1%
0%
1%
0%
1%
13%
39%
28%
88%
94%
76%
20%
10%
18%
6%
18%
17%
4%
3%
95%
96%
97%
100%
99%
99%
87%
85%
H
2
O

Figure 7.10: Reaction pathway diagram for n-butanol oxidation in the opposed-ow
diusion ame at T=858 K (italicized text), T=1170 K (bold text), and T=1520 K
(normal text).
Figure 7.12 displays the experimental and model predicted laminar burning velocities
at a number of equivalence ratios. The experimental data obtained in this study indicates
that the burning velocity increases between o=0.8 and o=1.1, which corresponds to a
maximum burning velocity of 47.7 cm/s, and then decreases at higher equivalence ratios.
There is good agreement between the presently obtained data and the data obtained by
Roberts nearly 50 years ago. The model well predicts the change in laminar burning
velocity as the equivalence ratio is increased from 0.7 to 1.4. The maximum burning
velocity predicted by the model is 45.8 cm/s at o=1.1, which is 1.9 cm/sec less than the
experimentally obtained value. Thus, it is observed that the proposed model performs
well in premixed ame environments.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 127



-0.06 -0.04 -0.02 0 0.02 0.04
C4H9OH+H=>H2+BC4H8OH
C4H9OH(+M)<=>C4H8+H2O(+M)
C4H9OH=>CH3+CC3H6OH
C4H9OH=>C2H5+C2H4OH
C4H9OH+H=>H2+CC4H8OH
C4H9OH+H=>H2+DC4H8OH
C2H4+H<=>C2H3+H2
C2H4+CH3<=>C2H3+CH4
C3H6+H<=>AC3H5+H2
2CH3(+M)<=>C2H6(+M)
AC3H5+H(+M)<=>C3H6(+M)
Normalised Sensitivity Coefficient
858K
1170 K

Figure 7.11: Sensitivity of n-butanol concentration to select reactions in the atmospheric
opposed-ow diusion ame (6% C
4
H
9
OH, 42% O
2
).
7.5 Conclusions
In the JSR and opposed-ow diusion ame, the proposed model indicates that H-
abstraction is the major pathway for n-butanol consumption, followed by -scission of
the resulting fuel radicals. These ndings dier from previous studies on n-butanol con-
sumption because of the peculiarity of each experimental apparatus used. In a pyrolysis
study of n-butanol, Barnard suggested that ssion at the C
3
H
7
-CH
2
OH bond to produce
the n-propyl radical and hydroxymethyl radical was dominant [9]. Barnards ndings are
consistent with what one would expect from a pyrolysis study where oxygen is absent,
hence unimolecular dissociation dominates. McEnally and Pfeerle studied coowing
laminar non-premixed methane ames doped with n-butanol, and concluded that uni-
molecular dissociation dominated over H-atom abstraction [11]. Their ndings are logical
since it applies to a doped coow ame in the centerline region, where the temperatures
are very high (1300 K), the fuel concentration is at a maximum, and the concentra-
tions of radical species are at a minimum. It is not clear if the proposed unimolecular
decomposition mechanism would dominate in coow ames of pure n-butanol (i.e. un-
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 128
25
30
35
40
45
50
0.7 0.9 1.1 1.3 1.5
Equivalence Ratio
L
a
m
i
n
a
r

B
u
r
n
i
n
g

V
e
l
o
c
i
t
y

(
c
m
/
s
)
Current Study
Roberts (1959)
Numerical

Figure 7.12: Laminar burning velocities of n-butanol/air mixtures, T=350 K, P=90 kPa.
doped ames). H-abstraction dominates in the experimental congurations presented
herein due to the nature of the combustion processes. The JSR is a premixed apparatus
so there is a rapid formation of radical species, which contribute to H-atom abstraction
reactions being dominant. H-atom abstraction dominates in the opposed ow diusion
ame because radical species generated near the ame front are able to diuse into the
fuel stream to consume the fuel. Unimolecular decomposition is not signicant even at
regions where radical species concentrations are small because the temperature of the
fuel stream is low at these points.
In this study, experimental data for n-butanol oxidation in a JSR (101.3 kPA and 1013
kPa), opposed-ow diusion ame, and premixed laminar ame have been compared to
a chemical kinetic mechanism. The experimental data in the JSR and opposed-ow dif-
fusion ame indicate that the most abundant measured species were CO, CO
2
, H
2
, H
2
O,
CH
4
, and C
2
H
4
. Appreciable quantities of the oxygenated species, butanal, acetaldehyde,
and formaldehyde, were also detected. The proposed model provides good qualitative
agreement with the data obtained across the three experimental congurations. Good
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 129
quantitative agreement is also observed for a number of species, however the potential
for improvement exists.
7.6 Recommendations
Both the proposed model and the experiments suggest that enols (i.e., R
1
R
2
C=CH(OH))
are minor species produced during the oxidation of n-butanol in both the JSR and the
opposed-ow diusion ame. Their predicted maximum concentration at low ppm levels
(<5 ppm) in the opposed diusion ame and sub-ppm levels in the JSR suggest that their
chemistry is of little importance for modeling n-butanol oxidation in these experiments.
However, it is well known that enols rapidly tautomerize into aldehydes, so it is possible
that the experimental data presented here did not accurately measure enol concentrations
in the ame. Other studies have identied enols as important species in the combustion
of oxygenates [34, 12] and hydrocarbons [35, 36]. Recent ab initio rate calculations by
Simmie et. al [37] suggest that butenol (i.e., 1-buten-1-ol) should be as abundant as
butanal during the combustion of n-butanol. An improved n-butanol chemical kinetic
mechanism has been developed by the authors, which includes reaction pathways and
rate estimates for the production and destruction of enols [38], but additional experi-
mental data is required to validate the mechanism. In summary, an improved n-butanol
mechanism would benet from the following:
1. Fundamental studies, experimental and theoretical, on n-butanol H-abstraction
reaction rates across a wide range of temperatures.
2. Fundamental reaction rate studies on simple and complex n-butanol unimolecular
decomposition reactions at intermediate and high temperatures.
3. Well validated chemical kinetic sub-mechanisms for enols and aldehydes.
It was mentioned previously that ABE is a biofuel that has a potential for use in
combustion applications; however, there are currently no systems designed to use this
fuel. Future research should be directed towards enabling the use of ABE biofuel in
combustion applications through detailed analysis of the fuels combustion properties.
In addition, the studies should focus on identifying and mitigating pollutant formation
associated with ABE combustion.
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 130
Detailed chemical kinetic models have been developed for butanol, ethanol, and ace-
tone [39], but none are capable of dealing with ABE mixtures. Research should be
directed towards developing a validated chemical kinetic model for ABE mixtures. The
validated kinetic model will enable engine designers to optimize engine systems for ABE
combustion. The species measured in dierent experimental setups will provide a better
understanding of pollutant formation under varying operating conditions (i.e. tempera-
ture, pressure, air-fuel ratios, etc.).
In addition to detailed chemical kinetic combustion models, future research should be
directed towards developing skeletal mechanisms for butanol and ABE biofuels. These
simplied mechanisms are advantageous because they require less computational power
to solve, while still allowing the prediction of combustion product species. In order to
develop these simplied models, one must rst start with the aforementioned detailed
chemical kinetic model and set of experimental data. A reduction methodology, such as
the direction relation graph method, can then be used to create a skeletal mechanism
[40]. Thus, the detailed chemical kinetic mechanism provides a fundamental scientic
basis for biofuel combustion while the skeletal mechanism is used in the modeling of
practical combustion systems.
Supplemental Material
The journal publication [19] corresponding to this study includes the following supple-
mental material. Figures corresponding to data and modeling predictions are also avail-
able in Appendix B:
1. Raw experimental data, modeling predictions, and corresponding graphs for n-
butanol in the JSR at 10 atm. (.XLS format)
2. Raw experimental data, modeling predictions, and corresponding graphs for n-
butanol in the JSR at 1 atm. (.XLS format)
3. Raw experimental data, modeling predictions, and corresponding graphs for n-
butanol in the opposed-ow diusion ame. (.XLS format)
4. Raw experimental data, modeling predictions, and corresponding graphs for n-
butane in the opposed-ow diusion ame. (.XLS format)
Chapter 7. Chemical Kinetic Modeling of Butanol Combustion 131
5. The proposed n-butanol chemical kinetic mechanism in CHEMKIN format. (.INP
format)
6. The proposed n-butanol thermodynamic datale in CHEMKIN format. (.DAT
format)
7. The proposed n-butanol transport datale in CHEMKIN format. (.DAT format)
Literature Cited
[1] M. Gautam, D. Martin, and D. Carder, Emissions characteristics of higher alco-
hol/gasoline blends, Proceedings of the Institution for Mechanical Engineers, vol.
214, pp. 165182, 2000.
[2] M. Gautam and D. Martin, Combustion characteristics of higher-alcohol/gasoline
blends, Proceedings of the Institution for Mechanical Engineers, vol. 214, pp. 497
511, 2000.
[3] R. Ricee, A. Sanyal, A. Elrod, and R. Bata, Exhaust-gas emissions of butanol,
ethanol, and methanol-gasoline blends, Journal of Engineering for Gas Turbines
and Power, vol. 113, no. 3, pp. 377381, July 1991.
[4] F. N. Alasfour, Butanol - a single-cylinder engine study: engine performance,
International Journal of Energy Research, vol. 21, pp. 2130, 1997.
[5] F. Alasfour, Butanol - a single-cylinder engine study: availability analysis, Applied
Thermal Engineering, vol. 6, pp. 537549, 1997.
[6] Y. Yacoub, R. Bata, and M. Gautam, The performance and emission characteristics
of C1-C5 alcohol-gasoline blends with matched oxygen content in a single-cylinder
spark ignition engine, Proceedings of the Institution of Mechanical Engineers Part
A - Journal of Power and Energy, vol. 212, no. A5, pp. 363379, 1998.
[7] G. Miller, J. Smith, and J. Workman, Engine performance using butanol fuel
blends, Transactions of the American Society of Automotive Engineers, vol. 24,
pp. 538540, 1981.
[8] T. Wallner, S. A. Miers, and S. McConnell, A comparison of ethanol and butanol as
oxygenates using a direct-injection, spark-ignition engine, Journal of Engineering
for Gas Turbines and Power, vol. 131, no. 3, May 2009.
132
Literature Cited 133
[9] J. Barnard, The pyrolysis of normal-butanol, Transactions of the Faraday Society,
vol. 53, no. 11, pp. 14231430, 1957.
[10] A. Roberts, The burning velocities of the alcohols, Journal of the Imperial College
of Chemical Engineering Society, vol. 12, pp. 5873, 1959.
[11] C. McEnally and L. Pfeerle, Fuel decomposition and hydrocarbon growth pro-
cesses for oxygenated hydrocarbons: butyl alcohols, Proceedings of the Combustion
Institute, vol. 30, no. Part 1, pp. 13631370, 2005, 30th International Symposium
on Combustion, Chicago, IL, July 25-30, 2004.
[12] B. Yang, P. Osswald, Y. Li, J. Wang, L. Wei, Z. Tian, F. Qi, and K. Kohse-
Hoeinghaus, Identication of combustion intermediates in isomeric fuel-rich pre-
mixed butanol-oxygen ames at low pressure, Combustion and Flame, vol. 148,
no. 4, pp. 198209, March 2007.
[13] J. T. Moss, A. M. Berkowitz, M. A. Oehlschlaeger, J. Biet, V. Warth, P.-A. Glaude,
and F. Battin-Leclerc, An experimental and kinetic modeling study of the oxidation
of the four isomers of butanol, Journal of Physical Chemistry A, vol. 112, no. 43,
pp. 10 84310 855, October 2008.
[14] P. Dagaut and C. Togbe, Oxidation kinetics of butanol-gasoline surrogate mixtures
in a jet-stirred reactor: Experimental and modeling study, Fuel, vol. 87, no. 15-16,
pp. 33133321, November 2008.
[15] C. McEnally, U. Koylu, L. Pfeerle, and D. Rosner, Soot volume fraction and
temperature measurements in laminar nonpremixed ames using thermocouples,
Combustion and Flame, vol. 109, no. 4, pp. 701720, June 1997.
[16] P. Dagaut and M. Cathonnet, The ignition, oxidation, and combustion of kerosene:
A review of experimental and kinetic modeling, Progress in Energy and Combustion
Science, vol. 32, no. 1, pp. 4892, 2006.
[17] P. Dagaut and S. Gail, Chemical kinetic study of the eect of a biofuel additive
on jet-a1 combustion, Journal of Physical Chemistry A, vol. 111, no. 19, pp. 3992
4000, May 2007.
[18] R. J. Kee, F. M. Rupley, J. A. Miller, M. E. Coltrin, J. F. Grcar, E. Meeks, H. K.
Moat, A. E. Lutz, G. Dixon-Lewis, M. D. Smooke, J. Warnatz, G. H. Evans,
Literature Cited 134
R. S. Larson, R. E. Mitchell, L. R. Petzold, W. C. Reynolds, M. Caracotsios, W. E.
Stewart, P. Glarborg, C. Wang, C. L. McLellan, O. Adigun, W. G. Houf, C. P. Chou,
S. F. Miller, P. Ho, P. D. Young, D. J. Young, D. W. Hodgson, M. V. Petrova, and
K. V. Puduppakkam. (2006) CHEMKIN release 4.1. San Diego, California.
[19] S. M. Sarathy, M. J. Thomson, C. Togbe, P. Dagaut, F. Halter, and C. Mounaim-
Rousselle, An experimental and kinetic modeling study of n-butanol combustion,
Combustion and Flame, vol. 156, no. 4, pp. 852864, April 2009.
[20] T. L. Cong and P. Dagaut, Kinetics of natural gas, natural gas/syngas mixtures
oxidation and eect of burnt gas recirculation: Experimental and detailed modeling,
in Proc. ASME Turbo Expo 2007: Power for Land, Sea and Air, vol. GT2007-27146,
Montr
`
Eal, Canada, May 2007, pp. 19.
[21] P. Dagaut, On the kinetics of hydrocarbons oxidation from natural gas to kerosene
and diesel fuel, Physical Chemistry Chemical Physics, vol. 4, no. 11, pp. 20792094,
2002.
[22] F. Battin-LeClerc, R. Bounaceur, G. Come, and et. al. (2004) Exgas-alkanes, a
software for the automatic generation of mechanisms for the oxidation of alkanes.
CNRS-DCPR.
[23] H. Curran, Rate constant estimation for C1 to C4 alkyl and alkoxyl radical decom-
position, International Journal of Chemical Kinetics, vol. 38, no. 4, pp. 250275,
April 2006.
[24] N. Marinov, A detailed chemical kinetic model for high temperature ethanol ox-
idation, International Journal of Chemical Kinetics, vol. 31, no. 3, pp. 183220,
March 1999.
[25] J. Park, Z. Xu, and M. Lin, Thermal decomposition of ethanol. ii. a computational
study of the kinetics and mechanism for the H+C2H5OH reaction, Journal of
Chemical Physics, vol. 118, no. 22, pp. 99909996, June 2003.
[26] C. Muller, V. Michel, G. Scacchi, and G. Come, THERGAS - a computer-program
for the evaluation of thermochemical data of molecules and free-radicals in the gas-
phase, Journal de Chimie Physique et de Physico-Chimie Biologique, vol. 92, no. 5,
pp. 11541178, May 1995.
Literature Cited 135
[27] S. Benson, Thermochemical Kinetics, 2nd edition. Wiley, New York, 1976.
[28] S. Gail, M. Thomson, S. Sarathy, and S. Syed, A wide-ranging kinetic modeling
study of methyl butanoate combustion, Proceedings of the Combustion Institute,
vol. 31, pp. 305311, 2007.
[29] L. Tee, S. Gotoh, and W. Stewart, Molecular parameters for normal uids - lennard-
jones 12-6 potential, Industrial & Engineering Chemistry Fundamentals, vol. 5,
no. 3, p. 356, 1966.
[30] H. Wang and M. Frenklach, Transport properties of polycyclic aromatic hydrocar-
bons for ame modeling, Combustion and Flame, vol. 96, no. 1-2, pp. 163170,
January 1994.
[31] P. Linstrom and W. Mallard. (2005) NIST chemistry webbook, NIST standard
reference database number 69. National Institute of Standards and Technology,.
Gaithersburg, Maryland, 20899. [Online]. Available: http://webbook.nist.gov
[32] D. R. Lide, Ed., CRC Handbook of Chemistry and Physics, 87th Edition. Boca Ra-
ton, FL: Taylor and Francis, 2007. [Online]. Available: http:/www.hbcpnetbase.com
[33] A. McClellan, Tables of Experimental Dipole Moments. San Francisco: Freeman,
1963.
[34] Y. Li, L. Wei, Z. Tian, B. Yang, J. Wang, T. Zhang, and F. Qi, A comprehensive
experimental study of low-pressure premixed C3-oxygenated hydrocarbon ames
with tunable synchrotron photoionization, Combustion and Flame, vol. 152, no. 3,
pp. 336359, February 2008.
[35] C. Taatjes, N. Hansen, J. Miller, T. Cool, J. Wang, P. Westmoreland, M. Law,
T. Kasper, and K. Kohse-Hoinghaus, Combustion chemistry of enols: Possible
ethenol precursors in ames, Journal of Physical Chemistry A, vol. 110, no. 9, pp.
32543260, March 2006.
[36] C. Taatjes, N. Hansen, A. McIlroy, J. Miller, J. Senosiain, S. Klippenstein, F. Qi,
L. Sheng, Y. Zhang, T. Cool, J. Wang, P. Westmoreland, M. Law, T. Kasper, and
K. Kohse-Hoinghaus, Enols are common intermediates in hydrocarbon oxidation,
Science, vol. 308, no. 5730, pp. 18871889, June 2005.
Literature Cited 136
[37] J. M. Simmie and H. J. Curran, Energy barriers for the addition of h, CH3, and
C2H5 to CH2=CHX [X = H, CH3, OH] and for h-atom addition to RCH=O [R = H,
CH3, C2H5, n-C3H7]: Implications for the gas-phase chemistry of enols, Journal
of Physical Chemistry A, vol. 113, no. 27, pp. 78347845, July 2009.
[38] G. Black, H. Curran, S. Pichon, J. Simmie, and V. Zhukov, Bio-butanol: Combus-
tion properties and detailed chemical kinetic model, Combustion and Flame, vol.
in press, p. doi:10.1016/j.combustame.2009.07.007, 2009.
[39] S. Pichon, G. Black, N. Chaumeix, M. Yahyaoui, J. M. Simmie, H. J. Curran, and
R. Donohue, The combustion chemistry of a fuel tracer: Measured ame speeds and
ignition delays and a detailed chemical kinetic model for the oxidation of acetone,
Combustion and Flame, vol. 156, no. 2, pp. 494504, February 2009.
[40] T. Lu and C. K. Law, Toward accommodating realistic fuel chemistry in large-
scale computations, Progress in Energy and Combustion Science, vol. 35, no. 2, pp.
192215, April 2009.
Part III
Biodiesel
137
Chapter 8
Background
Biodiesel is a fuel comprised of fatty acid alkyl esters derived from vegetable oils or
animal fats and meeting the requirements of ASTM D6751 [1]. Biodiesel is of particular
interest because it can replace petroleum diesel for use in compression ignition engines.
Commercial biodiesel fuels are produced via transesterication of triglycerides extracted
from a variety of biolipid feedstocks including the following: virgin vegetable oil feedstocks
such as rapeseed, soybean, canola, mustard, palm oil, and sunower; waste vegetable oils;
animal fats such as beef tallow, chicken fat, lard and yellow grease; and non-edible oils
such as jatropha, neem oil, castor oil, tall oil, and microalgal oil [2, 3]. The advantages
of biodiesel include the following[4]:
It is a renewable fuel.
It displaces the use of petroleum diesel.
It has a low toxicity and high biodegradability.
It may help reduce greenhouse gas emissions.
It can reduce particulate matter (PM), carbon monoxide (CO), sulphur oxides
(SO
a
), and total hydrocarbon (THC) emissions.
It can be made locally from agricultural and/or recycled feedstocks.
Biodiesel refers to the pure fuel before blending with diesel fuel. Biodiesel blends are
denoted as, BXX with XX representing the percentage of biodiesel contained in the
blend (i.e., B20 is 20% biodiesel, 80% petroleum diesel) [5]. Biodiesel can be used in its
138
Chapter 8. Background 139
pure form (i.e., B100), or it can be blended with petroleum diesel, as is more common in
practice. A blend of 20% biodiesel (i.e., B20) is the most common in the United States
because it balances performance, cost, emissions, and handling with petroleum diesel
[4]. Biodiesel is typically comprised of a mixture of saturated and unsaturated fatty acid
alkyl esters (e.g., methyl esters, ethyl esters, etc.) with chain lengths ranging from 12
to 18 carbon atoms. The spectrum of fatty acid alkyl ester composition depends on the
feedstock oil and the alcohol (i.e., methanol or ethanol) used during production.
In order to optimize engine systems for biodiesel, designers require information about
the fundamental combustion chemistry of the fuel (e.g., chemical kinetic mechanisms).
Developing chemical kinetic mechanisms for biodiesel has been challenging due to the
large size of the fatty acid alkyl esters found in practical fuels and the added complex-
ity of varying chain length and degrees of unsaturation. Nevertheless, much progress has
been made in understanding biodiesel combustion chemistry. This chapter presents back-
ground information on biodiesel so that the reader understands the fuels complexity and
the need for combustion studies to better understand the eects on engine performance.
Biodiesel Sustainability
Biodiesel is considered the most sustainable amongst current food-based biofuels because
the energy needed for processing oils and fats into biofuel is small [6]. The most widely
used feedstock for biodiesel are soybean oil in the U.S. and canola/rapeseed oil in Canada
and Europe. A large number of studies have applied the LCA approach for assessing
the sustainability of biodiesel fuels compared to conventional petroleum diesel. Several
LCA review papers have reported that soybean methyl ester can oer reductions in
fossil energy use ranging from 50% to 93%, and reductions in GHG emissions ranging
from 16% to 63% [6, 7, 8]. However, some studies report substanial increases in GHG
emissions for some seed-based biodiesel fuels [9] The wide error margins around the exact
values for reductions in fossil energy use and GHG emissions are because of the dierent
assumptions, coproduct credit allocation methods, and other case-specic information
used in dierent studies.
Biodiesel has also been cited as potentially harmful to the environment and society.
Fargione et al. calculated that the conversion of tropical rainforests to cropland for palm
oil methyl ester results in carbon debts that may take hundred of years to repay [10]. As
palm oil is also used for food, its diversion towards biodiesel has societal impacts, and the
Chapter 8. Background 140
deforestation occurring for crop production threatens tropical biodiversity [11]. Oilseed
biodiesel fuels oer limited environmental benets because they harness only a small
portion of the available above-ground biomass, and thus they are disadvantaged from a
yield perspective [9]. Agricultural, environmental, and societal impacts can be largely
reduced by using crops that require less agricultural inputs (i.e., fertilizer, pesticide,
energy, and water), using marginal or low agricultural value land, and using non food-
based feedstock [6, 11].
Waste vegetable oils and animal fats are a technologically and economically viable
feedstock for biodiesel production [12, 13]. However, the limited quantity, high contami-
nant levels, and poor cold performance characteristics [14] limit the ability of these waste
feedstock to signicantly displace petroleum diesel. Nevertheless, such waste feedstock
technologies are considered sustainable alternatives to conventional oilseed biodiesel.
Recently, biodiesel fuel derived from microalgae has become the focus of research in-
terest. Microalgae are small photosynthetic organisms that convert sunlight and carbon
dioxide into biomass. Chisti [3] calculated that there is not enough land in the U.S. for
sustainable cropping of oilseeds to displace 50% of petroleum diesel consumption; how-
ever, only 1-3% of total U.S. cropping area would be needed to replace 50% of transport
fuel needs using microalgal biodiesel. The present state of the technology is technically
feasible, but substantial improvements in genetic and metabolic engineering are needed to
make microalgal biodiesel economically competitive at commercial scales [3]. Additional
research is also needed to understand the fuel properties (e.g., physical chemical proper-
ties, combustion properties, etc.) of microalgal biodiesel, since its chemical composition
is dierent than conventional vegetable oil and animal fat based biodiesels.
8.1 Biodiesel Fuel Chemistry
As previously mentioned, biodiesel is derived from fats and oils, which are comprised of
compounds termed triglycerides
1
. Triglycerides consist of one molecule of glycerol com-
bined with three molecules of fatty acid. A fatty acid is a hydrocarbon chain terminating
in a carboxyl group. If the three fatty acids in the triglyceride are identical, then it is
a simple triglyceride. However, most triglycerides are mixed, meaning they contain a
mixture of dierent fatty acids. Thus, the properties of mixed triglycerides, and the fats
1
Triglycerides are properly known as trialcylglycerols
Chapter 8. Background 141
and oils which they comprise, depend on the fatty acid content.
Most vegetable oils and animal fats consist of fatty acids containing 12 and 18 carbon
atoms. If a double bond exists between two carbon atoms, then the chain is described as
unsaturated because all available carbon valences for hydrogen are not satised. Unsat-
urated fatty acids with two or more double bonds are called polyunsaturated fatty acids,
and the extent of unsaturation is quantied by the iodine value.
There are a variety of feedstocks available for biodiesel production, but most can be
classied as vegetable oils, rst-use animal fats, or waste greases [15]. The molecular
structure and physical chemical properties of a given biodiesel are directly related to the
fatty acid composition of the fat or oil from which the biodiesel was derived [16]. Table 8.1
presents the fatty acid composition of several common feedstock for biodiesel production.
Biodiesel derived from highly unsaturated feedstock, such as canola or soybean oil, will be
equally unsaturated. Similarly, biodiesel derived from palm kernel or coconut oil is highly
saturated because the feedstock consists of primarily saturated fatty acids. Microalgal
oils used for biodiesel are rich in polyunsaturated fatty acids with four or more double
bonds and chain lengths exceeding 20 carbon atoms [3].
Studying the combustion chemistry of biodiesel fuels is challenging because the molec-
ular structure and physical chemical properties depend strongly on feedstock selection.
The development of detailed chemical kinetic mechanisms is dicult due to the numer-
ous possible reaction pathways for long chain fatty acid alkyl ester molecules. The added
complexity of varying chain length and degrees of unsaturation has led to the use of surro-
gate fuels of well characterized composition for chemical kinetic studies. The purpose of
using surrogates is to simplify the combustion mechanism by using a single fuel molecule
to represent the mixtures of alkyl esters, and by using smaller molecules to reduce the
number of possible chemical reactions.
8.1.1 Biodiesel Production
Producing biodiesel that meets the ASTM D6751 standards requires the feedstock oil
or fat to be processed, so that its viscosity is reduced. Several methods by which the
viscosity can be reduced are micro-emulsication, cracking, and transesterication. Ma
and Hanna have reviewed these processes [18]; however, only transesterication produces
a fuel that meets the ASTMs denition of biodiesel [1], wherein biodiesel is comprised
of long chain mono alkyl esters.
Chapter 8. Background 142
Table 8.1: Typical fatty acid composition (wt%) of various biodiesel feedstock
a
[17]
Lauric Myristic Palmitic Stearic Oleic Linoleic Linolenic
C
12
:0 C
14
:0 C
16
:0 C
18
:0 C
18
:1 C
18
:2 C
18
:3
Beef Tallow 0 3 24 19 43 3 1
Canola 0 0 4 2 62 22 10
Coconut 47 18 9 3 6 2 0
Palm 0 0 45 4 40 10 0
Palm kernel 48 16 8 3 15 2 0
Soybean 0 0 7 5 19 68 7
a
The fatty acid molecular structure is indicated by C
a
:y, where x is the carbon
number and y is the number of double bonds. Compositions may not add to
100% because trace fatty acids are not listed.
Transesterication is the process by which a triglyceride is converted to a mono alkyl
ester. It involves reacting the fat or oil with an alcohol, in the presence of a catalyst,
to produce glycerol and esters [19]. Methanol is the most commonly used alcohol, and
therefore the resulting biodiesel is often referred to as a mixture of FAME. Ethanol has
also been used [20], and in such cases the resulting biodiesel is comprised of fatty acid
ethyl esters. The present study focuses on biodiesel comprised of FAME moieties since
most biodiesel in production is of this type [16].
The transesterication reaction between a typical triglyceride and methanol is shown
in Figure 8.1. Three moles of methanol are required to react with each mole of triglyceride
in the oil or fat. The reaction is carried out in the presence of a catalyst to improve the
reaction rate [19]. The catalyst, alcohol, and vegetable oil are combined in a batch reactor
at approximately 65

C and stirred continuously. The duration of the reaction can range
from 1-6 hours, depending on the desired yield. Two immiscible layers are formed once
the reversible reaction reaches equilibrium. The lower layer is glycerol and the upper
layer contains FAME and unreacted feedstock. Many processes separate the glycerol and
conduct a second transesterication reaction to increase the yield. If the biodiesel is to
meet ASTM D6751 specications, it must undergo a series of separation processes for
the removal of alcohol, catalyst, water, soaps, glycerol, and unreacted triglycerides and
free fatty acids [21]. The separated crude glycerol can be further puried and used in
other applications (e.g., cosmetics, pharmaceuticals, etc.).
Chapter 8. Background 143
Figure 8.1: Transesterication Reaction
8.2 Biodiesel Fuel Properties
Biodiesel must have physical-chemical properties similar to that of diesel fuel in order
to successfully operate in a compression ignition engine. As mentioned previously, the
properties of biodiesel depend on the molecular structure of the FAME which comprise
it (i.e., their chain length and degree of unsaturation). The important fuel properties
which are inuenced by the FAME prole are exhaust emissions, LHV, ignition quality
(i.e., cetane number), viscosity, oxidative stability, lubricity, and cold ow characteristics.
This section summarizes the important combustion related properties of several biodiesel
fuels, pure FAME, and standard low sulfur diesel
2
.
The LHV is a measure of the fuels energy density. Diesel engines are capable of
accepting a variation in heating values, so there is no specied LHV in the ASTM D 975
Standard. However, it is benecial for biodiesel to have a volumetric LHV similar to that
of standard diesel so that dierences in fuel economy (i.e., L/100 km) are not experienced
[19]. The volumetric LHV of biodiesel is slightly lower than that of standard diesel, so
fuel economy would be lower [8]. It should be noted that some studies have reported an
improvement in engine eciency when using biofuels, which osets the lower volumetric
heating value, and results in no net change in fuel economy [23]. For pure FAME, the
2
Information on viscosity, oxidative stability, lubricity, and cold ow characteristics is available in the
literature [21, 16, 8, 22]
Chapter 8. Background 144
LHV increases with chain length [16].
The cetane number is a measure of the fuels ignition delay. Diesel combustion requires
the fuel to self-ignite as it is sprayed into the compressed cylinder gas. The self-ignition
leads to the characteristic diesel knock, wherein an explosion of premixed air and fuel
causes a rapid heat release and pressure rise. The magnitude of the explosion can be
decreased by shortening the ignition delay time. Higher cetane numbers result in shorter
ignition delay times, and therefore better engine operation. The ASTM D 975 minimum
acceptable cetane value is 40. Table 8.2 presents cetane numbers for common biodiesel
fuels and pure FAME. The iodine value is also shown for biodiesel fuels to indicate
the degree unsaturation. Biodiesel fuels have higher cetane numbers when compared to
standard diesel [8, 21, 22]. Similar to hydrocarbon compounds, the cetane number of
pure FAME decreases with increasing unsaturation and increases with increasing chain
length [16]. Soybean methyl ester and canola methyl ester have lower cetane numbers
than palm oil methyl ester because of the higher unsaturated fatty acid content in the
former [22]. Microalgal oils, which are rich in polyunsaturated fats with four or more
double bonds, have high iodine values and are likely to have depressed cetane numbers.
Current European biodiesel standards limit the iodine value and the concentration of
polyunsaturated FAME, so microalgal biodiesel must undergo hydrogenation
3
to meet
current fuel standards [3].
8.3 Biodiesel Exhaust Emissions
There have been a number of studies on the use of biodiesel in CI engines. The goal of
these studies was to determine the eect of using biodiesel and diesel-biodiesel blends on
the emissions, fuel economy, and operation of the engine. Comprehensive review articles
on the use of biodiesel in CI engines and it eects on engine performance and exhaust
emissions have been published recently [8, 23].
The eect of biodiesel on chemical emissions depends on the specic pollutant of
concern, the type of engine used, the engine speed and load, and the biodiesel FAME
composition. The present study is mainly concerned with the role of FAME composition
because detailed chemical kinetic mechanisms can be used to elucidate any chemistry
related eects on combustion emissions. Therefore, this section briey summarizes the
3
Hydrogenations will saturate the double bonds and decrease the iodine value.
Chapter 8. Background 145
Table 8.2: Properties of common biodiesel fuels and pure FAME
a
Biodiesel feedstock or pure FAME Cetane number Iodine value
Beef Tallow 75 33-47
Palm Oil 61 49-55
Rapeseed/Canola Oil 55 110-126
Soybean Oil 49 118-139
Caprylic FAME (C
8
:0) 33.6 -
Lauric FAME (C
12
:0) 61.4 -
Myristic FAME (C
14
:0) 66.2 -
Palmitic FAME (C
16
:0) 74.5 -
Stearic FAME (C
18
:0) 86.9 -
Oleic FAME (C
18
:1) 55 -
Linoleic FAME (C
18
:2) 42.2 -
a
Cetane numbers for biodiesel fuels are from [22, 21] and pure
FAME from [16]. Iodine values are from [24].
eects of biodiesel on diesel engine emissions with a particular focus on the role of FAME
composition
4
.
Table 8.3 presents the percentage of research publications that report increases, simi-
larities, or decreases in emissions when using biodiesel and diesel fuels [23]. The majority
of studies nd that the use of biodiesel reduces the emissions of THC, CO, and PM, and
increases the emissions of NO
a
. The eect of biodiesel on the emissions of oxygenated
compounds is uncertain since research studies have shown both increases, decreases, and
insignicant dierences compared to petroleum diesel.
8.3.1 CO, THC, and Oxygenate Emissions
The emissions of THC decreases when biodiesel is used [23]. Generally, the feedstock used
for biodiesel production does not aect THC emissions. However, research on pure FAME
indicates that THC emissions decrease with increasing fatty acid chain length [25, 26]
and decreasing unsaturation [25]. The most widely accepted reason for the decrease in
4
The information presented is from a comprehensive review discussing the eect of biodiesel on diesel
engine emissions by Lapuerta et al. [23]. The reader is directed to the original article for a more detailed
discussion.
Chapter 8. Background 146
Table 8.3: Percent of publications that report changes in emissions for biodiesel [23].
Increase Same Decrease
NO
a
85 10 5
PM 3 2 95
THC 1 4 95
CO 2 8 90
THC when compared to diesel fuel is that the oxygen content and higher cetane number
of biodiesel leads to a more complete and cleaner combustion. Rakapoulos has shown
that THC emissions decrease as the oxygen content in the cylinder increases, either by
enriching the oxygen content of the fuel or the air [27]. Increasing the cetane number
also helps reduce THC emissions, and this explains why longer chain and fully saturated
FAME have lower THC emissions.
CO emissions decrease when biodiesel is used [23]. For pure FAME, CO emissions
decrease with increasing chain length [16] and decreasing unsaturation [25, 28]. The de-
crease in emissions when compared to petroleum diesel is because of the higher oxygen
content and cetane number of biodiesel, which promote complete combustion. Cetane
numbers increase with increasing chain length and decreasing unsaturation, so this ex-
plains the decreases in CO emissions observed in longer chain and saturated FAME.
It is widely believed that biodiesel would lead to greater emissions of oxygenated
compounds, such as aldehydes and ketones, because of the fuel bound oxygen in FAME.
However, engine studies have shown varying results, and there is no conclusive evidence
on this matter [23]. One study that measured increased acrolein emissions in several
biodiesel fuels attributed this to the glycerol content of biodiesel [29]. The argument
appears valid since glycerol combustion is shown to produce high levels of acetaldehyde
and acrolein [30, 31].
8.3.2 PM and NO

emissions
It is nearly unanimous that PM emissions decrease when biodiesel is used [23]. The eects
of biodiesel feedstock and FAME molecular structure on PM emission is uncertain. The
EPA reported that PM emissions were lower for beef tallow methyl ester than soybean
methyl ester [28], which suggests that higher degrees of unsaturation may increase PM
emissions. However, studies on pure FAME have shown no correlation between FAME
Chapter 8. Background 147
molecular structure (i.e., chain length and unsaturation) and PM emissions [16, 25]. The
main factor aecting PM emissions is the oxygen content of the fuel, which is constant
for various biodiesel fuels and FAME [25]. There are a number of reasons explaining
the reductions in PM observed when using biodiesel. Biodiesel has no aromatic com-
pounds, so its use displaces the highly sooting aromatic compounds typically found in
petroleum diesel fuel. In additon, the oxygen atoms in the FAME bonds to carbon atoms,
and therefore prevents carbon atoms from participating in soot growth reactions. The
oxidation of FAME and other oxygenated intermediates also forms OH radicals, which
readily attack unsaturated hydrocarbons and prevent their participation in soot growth
reactions. The FAME combustion chemistry study presented in this dissertation oers
additional insights into the role of the ester moiety in reducing soot.
The use of biodiesel in diesel engines leads to a slight increase in NO
a
emissions
[23]. The EPA reported that NO
a
emissions were lower for beef tallow methyl ester
than soybean methyl ester [28], which suggests that higher degrees of unsaturation may
increase NO
a
emissions. For pure FAME, NO
a
emissions increase with decreasing chain
length and increasing unsaturation [16, 25].
Explanations for the observed increase in NO
A
in biodiesel are open for speculation.
The three prevailing mechanisms explaining NO
a
formation in combustion engines are: 1.
Fuel, 2. Thermal (i.e., Zeldovich mechanism), and 3. Prompt (i.e., Fenimore mechanism).
Fuel NO
a
is formed when nitrogen compounds xated in the fuel are oxidized. This is
not a concern for biodiesel since the fuel does not contain any chemically bound nitrogen.
Thermal NO
a
is a result of high temperature dissociation and chain reaction of elemen-
tal nitrogen and oxygen in the post-combustion regime. The reactions in this mechanism
are shown in Equations 8.1, 8.2, and 8.3. The reactions are highly temperature depen-
dent, such that a decrease in combustion temperature will decrease NO
a
formation [32].
Tat and Van Gerpen [33] have shown that biodiesel has a higher isentropic bulk modulus
of compressibility than conventional diesel, which causes an inadvertent advance in fuel
injection time in a pump-line-nozzle fuel injector. The advance in fuel injection timing
increases the ignition delay, and therefore increases thermal NO
a
formation. Boehman
and coworkers [34] have shown that the bulk modulus of compressibility increases with
increasing iodine value, so this may explain the higher NO
a
emissions observed for un-
saturated FAME. This injection related phenomenon is the most widely cited reason for
biodiesels increased NO
a
formation [23]. The higher NO
a
emissions observed for shorter
chain and unsaturated FAME suggests that the lower cetane numbers of these compounds
Chapter 8. Background 148
leads to longer ignition delay times, and therefore more NO
a
formation.
Some experiments have shown that NO
a
emissions increase for biodiesel even if the
injection timing is matched exactly with petroleum diesel; and therefore, other mecha-
nisms than advanced injection time are proposed. Mueller et al. [35] recently conducted
experiments to assess the various proposed mechanisms. The results suggest that NO
a
emissions increases because of advances in combustion phasing that lead to higher in-
cylinder temperatures and longer residence times, and lower radiative heat losses which
lead to higher ame temperatures. Therefore, the increased NO
a
emissions appear to be
largely attributed to the thermal NO
a
mechanism.
C +`
2

`C + ` (8.1)
` +C
2

`C + C (8.2)
` +CH

`C +H (8.3)
Research studies have not eliminated the possibility of prompt NO
a
routes being the
cause of higher biodiesel NO
a
emissions. The prompt mechanism of NO
a
formation is
highly dependent on hydrocarbon radical intermediates formed during combustion, so it
is possible that chemical kinetic eects play a role [23, 35]. Garner et al. [36] postulate
that unsaturated FAME in biodiesel lead to higher acetylene levels that contribute to
prompt NO
a
formation. If such a mechanism is real, then detailed chemical kinetic models
for FAME can help to elucidate the role of molecular structure on NO
a
formation.
Literature Cited
[1] ASTM, ASTM D 6751 specication for biodiesel fuel blend stock (B100) for
middle distillate fuels, in ASTM Book of Standards. ASTM, 2003.
[2] A. Demirbas, Importance of biodiesel as transportation fuel, Energy Policy,
vol. 35, no. 9, pp. 46614670, September 2007.
[3] Y. Chisti, Biodiesel from microalgae, Biotechnology Advances, vol. 25, no. 3, pp.
294306, May-June 2007.
[4] Anonymous, Biodiesel handling and use guidelines, fourth edition, U.S. Depart-
ment of Energy, Tech. Rep., 2009.
[5] (2006, February) Ocial site of the national biodiesel board. [Online]. Available:
http://www.biodiesel.org/
[6] J. Hill, E. Nelson, D. Tilman, S. Polasky, and D. Tiany, Environmental, economic,
and energetic costs and benets of biodiesel and ethanol biofuels, Proceedings of
the National Academy of Sciences, vol. 103, pp. 11 20611 210, 2006.
[7] H. Huo, M. Wang, C. Bloyd, and V. Putsche, Life-cycle assessment of energy use
and greenhouse gas emissions of soybean-derived biodiesel and renewable fuels,
Environmental Science and Technology, vol. 43, no. 3, pp. 750756, February 2009.
[8] A. Agarwal, Biofuels (alcohols and biodiesel) applications as fuels for internal com-
bustion engines, Progress in Energy and Combustion Science, vol. 33, no. 3, pp.
233271, June 2007.
[9] E. Larson, A review of life-cycle analysis studies on liquid biofuel systems for the
transport sector, Energy for Sustainable Development, vol. 2, pp. 109126, 2006.
149
Literature Cited 150
[10] J. Fargione, J. Hill, D. Tilman, S. Polasky, and P. Hawthorne, Land clearing and
the biofuel carbon debt, Science, vol. 319, no. 5867, pp. 12351238, February 2008.
[11] L. P. Koh and J. Ghazoul, Biofuels, biodiversity, and people: Understanding the
conicts and nding opportunities, Biological Conservation, vol. 141, no. 10, pp.
24502460, 2008.
[12] Y. Zhang, M. Dube, D. McLean, and M. Kates, Biodiesel production from waste
cooking oil: 1. process design and technological assessment, Bioresource Technology,
vol. 89, no. 1, pp. 116, August 2003.
[13] Y. Zhang, M. Dube, D. McLean, and M. Kates, Biodiesel production from waste
cooking oil: 2. economic assessment and sensitivity analysis, Bioresource Technol-
ogy, vol. 90, no. 3, pp. 229240, December 2003.
[14] M. Canakci, The potential of restaurant waste lipids as biodiesel feedstocks, Biore-
source Technology, vol. 98, no. 1, pp. 183190, January 2007.
[15] R. Kotrba, Everything under the sun, Biodiesel Magazine, pp. 2934, February
2006.
[16] G. Knothe, Dependence of biodiesel fuel properties on the structure of fatty acid
alkyl esters, Fuel Processing Technology, vol. 86, no. 10, pp. 10591070, June 2005.
[17] T. C. of the Institute of Shortening and I. Edible Oils, Food fats and
oils, ninth edition, Institute of Shortening and Edible Oils, 1750 New
York Avenue, NW, Washington, DC, Tech. Rep., 2006. [Online]. Available:
http://www.iseo.org/foodfats.htm.
[18] F. Ma and M. A. Hanna, Biodiesel production: A review, Bioresource Technology,
vol. 70, no. 1, pp. 1 15, 1999.
[19] A. Ramadhas, S. Jayaraj, and C. Muraleedharan, Use of vegetable oils as i.c. engine
fuels - a review, Renewable Energy, vol. 29, no. 5, pp. 727 742, 2004.
[20] J. Encinar, J. Gonzalez, J. Rodriguez, and A. Tejedor, Biodiesel fuels from veg-
etable oils: Transesterication of cynara cardunculus l. oils with ethanol, Energy
& Fuels, vol. 16, no. 2, pp. 443 450, 2002.
Literature Cited 151
[21] J. Kinast, Production of biodiesels from multiple feedstocks and properties of
biodiesels and biodiesel/diesel blends, National Renewable Energy Laboratory,
Tech. Rep., 2003.
[22] M. Ramos, C. Fernandez, A. Casas, L. Rodriguez, and A. Perez, Inuence of fatty
acid composition of raw materials on biodiesel properties, Bioresource Technology,
vol. 100, no. 1, pp. 261268, January 2009.
[23] M. Lapuerta, O. Armas, and J. Rodriguez-Fernandez, Eect of biodiesel fuels on
diesel engine emissions, Progress in Energy and Combustion Science, vol. 34, no. 2,
pp. 198223, April 2008.
[24] D. R. Lide, Ed., CRC Handbook of Chemistry and Physics, 87th Edition. Boca Ra-
ton, FL: Taylor and Francis, 2007. [Online]. Available: http:/www.hbcpnetbase.com
[25] M. Graboski, R. McCormick, T. Allerman, and A. Herring, The eect of biodiesel
composition on engine emissions from a DDC series 60 diesel engine, National
Renewable Energy Laboratory, Tech. Rep., 2003.
[26] G. Knothe, Dependence of biodiesel fuel properties on the structure of fatty acid
alkyl esters, Fuel Processing Technology, vol. 86, no. 10, pp. 10591070, June 2005.
[27] C. Rakopoulos, D. Hountalas, T. Zannis, and Y. Levendis, Operational and envi-
ronmental evaluation of diesel engines burning oxygen-enriched intake air or oxygen-
enriched fuels: a review, SAE, no. SAE paper 2004-01-2924., 2004.
[28] Anonymous, Draft Technical Report EPA420-P-02-001, United States Environ-
mental Protection Agency, Tech. Rep., 2002.
[29] M. Graboski, J. Ross, and R. McCormick, Transient emissions from no. 2 diesel
and biodiesel blends in a ddc series 60 engine, SAE Special Publications, no. 1179,
p. 55, 1996.
[30] Y. Stein, M. Antal, and M. Jones, A study of the gas-phase pyrolysis of glycerol,
Journal of Analytical and Applied Pyrolysis, vol. 4, no. 4, pp. 283296, 1983.
[31] R. Mills, Flame structure of puried-glycerol/air combustion in a wick-burner,
Masters thesis, University of Toronto, 2009.
Literature Cited 152
[32] M. Hess, M. Haas, T. Foglia, and W. Marmer, Eect of antioxidant addition on
nox emissions from biodiesel, Energy & Fuels, vol. 19, no. 4, pp. 17491754, 2005.
[33] M. E. Tat and J. H. Van Gerpen, Eect of temperature and pressure on the speed of
sound and isentropic bulk modulus of mixtures of biodiesel and diesel fuel, JAOCS,
Journal of the American Oil Chemists Society, vol. 80, no. 11, pp. 1127 1130, 2003.
[34] A. L. Boehman, D. Morris, J. Szybist, and E. Esen, The impact of the bulk modulus
of diesel fuels on fuel injection timing, Energy & Fuels, vol. 18, no. 6, pp. 1877
1882, 2004.
[35] C. J. Mueller, A. L. Boehman, and G. C. Martin, An experimental investigation of
the origin of increased no
a
emissions when fueling a heavy-duty compression-ignition
engine with soy biodiesel, SAE International, vol. 2009-01-1792, 2009.
[36] S. Garner, R. Sivaramakrishnan, and K. Brezinsky, The high-pressure pyrolysis of
saturated and unsaturated c
7
hydrocarbons, Proceedings of the Combustion Insti-
tute, vol. 32, pp. 461467, 2009.
Chapter 9
An Experimental and Kinetic
Modeling Study of Biodiesel
Combustion
9.1 Introduction
Real biodiesel is a complex mixture of FAME with diering chain lengths and degrees of
unsaturation, so it is much simpler to study the combustion chemistry of pure FAME.
However, the numerous possible reaction pathways for long chain FAME would result in
extremely large detailed chemical kinetic mechanisms. Such mechanisms are cumbersome
to develop and computationally expensive to solve in even the simplest physical reactor
models. Furthermore, conducting fundamental combustion experiments using long chain
(i.e., high molecular weight) FAME is challenging because vaporisation is dicult.
In order to avoid the diculties associated with long chain FAME, surrogate fuels
with shorter chain lengths and known physical chemical properties are chosen for biodiesel
combustion chemistry studies. Using surrogate fuels simplies the chemical kinetic mech-
anism by reducing the number of possible chemical reactions, while still representing the
role of the molecular structure in combustion (i.e., the role of the methyl ester moiety and
the role of carbon-carbon double bonds). In addition, surrogates fuels are more volatile,
and therefore easier to work with experimentally.
Figure 9.1 displays typical biodiesel FAME and several proposed surrogates. The
surrogate fuels are structurally similar to actual biodiesel FAME, and all but one contain
153
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 154
the ester moiety. The chain length and degree of unsaturation varies in the surrogate
fuels, so the individual eects of each can be understood. The remainder of this section
discusses the recent advances in the chemical kinetic modeling of FAME surrogate fuels.
Particular attention is placed on the chemistry related eects of the ester moiety, carbon
chain length, and carbon-carbon double bonds during combustion.
O
O
Methyl cis-9-octadecenoate
O
O
Methyl hexadecanoate Methyl octadecanoate
O
O
Biodiesel FAME
O
O
Methyl butanoate
O
O
Methyl trans-2-butenoate
O
O
n-Hexadecane
Methyl decanoate
Surrogates for biodiesel
(Methyl oleate) (Methyl stearate) (Methyl palmitate)
(Methyl butyrate) (Methyl crotonate) (Methyl caprate) (Cetane)
Figure 9.1: Biodiesel FAME and their surrogates
9.2 Mechanisms for Short Chain Methyl Esters
Fisher and coworkers [1] were the rst to develop a detailed chemical kinetic for a
biodiesel surrogate. The authors developed a chemical kinetic mechanism for methyl
butanoate (MB), a fully saturated short chain FAME. It was chosen as a modeling
surrogate of biodiesel because it was thought to be large enough to allow fast RO
2
iso-
merization reactions, which are important for low-temperature chemical reactions that
control fuel auto-ignition in CI engines. While the mechanism was comprehensive, the
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 155
authors were unable to robustly validate the model due to limited experimental data on
MB combustion.
Gail et al. [2] were the rst to extensively validate a slightly modied version of
the Fisher mechanism for MB using experimental data from a JSR, an opposed-ow
diusion ame, and a ow reactor. The mechanism consisted of 295 chemical species and
1498 reactions. Recently, a number of studies have been conducted to further study the
combustion of MB and validate chemical kinetic mechanisms. Schwartz et al. [3] studied
MB combustion in co-ow ames of methane doped with MB. A number of experiments
in shock tubes and rapid compressions machines at various temperatures, pressures, and
equivalence ratios have been conducted to study the autoignition characteristics of MB
[4, 5, 6, 7, 8]. In addition, theoretical studies have performed ab initio calculations of
thermochemical properties [9, 10, 11, 12] and kinetic rate parameters [8, 13, 14] for MB.
The various studies on MB have revealed consistent conclusions. Firstly, quantum
calculations of thermochemical properties suggest that MB is a good surrogate fuel for
representing the thermochemistry of saturated long chain FAME. However, autoignition
and low temperature experimental data indicate that MB does not exhibit the cool ame
and negative temperature coecient (NTC) behaviour, which are signicant character-
istics of the longer chain FAME found in biodiesel. Figure 9.2 presents simulations of
methyl butanoate and methyl decanoate in a JSR at an initial fuel concentration of 1%, a
pressure of 1013.25 kPa, and a range of temperatures. Methyl decanoate displays typical
biodiesel NTC reactivity between 600 and 800 K, while methyl butanoate shows no reac-
tivity in this temperature range. Vaughan et al. [15] and Hadjiali et al. [16] also found
that the ignition delay time of MB did not match well with those of longer chain FAME.
Therefore, due to its short chain length, MB is not a suitable surrogate for understanding
the low temperature reactivity of biodiesel. However, it can serve as a starting point for
the development of mechanisms for larger molecules.
The studies also demonstrate the fate of the ester moiety during combustion. Figure
9.3 displays one high temperature combustion pathway leading to an important inter-
mediate during MB combustion, specically the methoxycarbonyl radical (CH
3
OCO),
which is produced when the ester group breaks away from the fatty acid. Subsequently,
the methoxycarbonyl radical primarily decays to form methyl radical and CO
2
. From
a soot reduction standpoint, this decarboxylation of the ester is not an ecient use of
fuel-bound oxygen because two oxygen atoms are bonded to one carbon atom. It would
be more ecient if the ester moiety led to the production of CO since each oxygen atom
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 156
1.E-05
1.E-04
1.E-03
1.E-02
500 600 700 800 900 1000
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
MB MD
Figure 9.2: Computed proles obtained from the oxidation of methyl decanoate and
methyl butanoate in a JSR at o=1.0, P=1013.25 kPa, t=1 s, 0.1% fuel mole fraction.
in the ester group would sequester one carbon atom from participating in the production
of soot. This conclusion is consistent with the ndings of other studies on methyl esters
and biodiesel [17, 18, 19]. Furthermore, Pepiots-Desjardins and coworkers [20] studied
the sooting tendency of various oxygenated (e.g., alcohols, esters, aldehydes, etc.), and
it was concluded that ester moieties are less ecient at reducing soot than alcohol and
aldehydes moieties.
The aforementioned decarboxylation of the ester moiety also has implications on the
production of oxygenated hydrocarbon compounds. It is widely believed that FAME
would lead to higher emissions of oxygenated hydrocarbons, such as aldehydes and ke-
tones, because of the oxygen atoms present in the molecule. The chemical kinetic mech-
anism of MB indicates that at high temperatures in a ame about one-third of the fuel
bound oxygen goes to form CO
2
directly and two-thirds forms oxygenated hydrocarbons.
The present study oers additional insights into the fate of the ester moiety during FAME
combustion.
Besides the ester moiety, the long chain FAME found in biodiesel are also distinguished
by their degree of unsaturation. To better understand the role of unsaturation, Sarathy
et al. [21] and Gail et al. [22] conducted combustion studies on methyl trans-2-butenoate
(MC), which is the monounsaturated counterpart of MB. The authors provided a chemical
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 157
C

O C H
3
O
C
O
O
O C
- CH
3
- CH
3
O
C H
3
CH
2
C H
2
C
O C H
3
O
C H
3
CH

C H
2
C
O C H
3
O
- C
3
H
6
- H
Figure 9.3: One combustion pathway of methyl butanoate that depicts the fate of the
ester moiety.
kinetic mechanism for MC validated against experimental data in a JSR and an opposed-
ow diusion ame. When compared to MB, the experimental data indicates that MC
combustion leads to higher levels of unsaturated hydrocarbon species (e.g., acetylene,
propyne, and propadiene), which have strong sooting tendencies. The chemical kinetic
mechanism for MC reveals that higher levels of unsaturated hydrocarbon species are
produced because the double bond is preserved during fuel decomposition, thus leading
to stable alkenes and alkynes, as shown in Figure 9.4. Such an analysis can explain why
some engine studies [23] saw the unsaturated soybean methyl esters forming more soot
than the saturated beef tallow methyl esters. Furthermore, if Garner et al.s hypothesis
[24] that higher acetylene levels increase NO
a
formation via the prompt NO
a
mechanism,
then the analysis of MC may explain why unsaturated FAME exhibit higher NO
a
in
engine studies.
9.3 Mechanisms for Long Chain Methyl Esters
The studies on C
4
FAME revealed that such short chain molecules are not suitable
surrogate fuels because they do not exhibit the low temperature autoignition properties
of long chain FAME. Therefore, chemical kinetic studies have shifted focus towards larger
molecules, in order to better represent the combustion properties of actual biodiesel. The
following is a discussion of the current progress in the modeling of larger surrogate fuels.
Dayma et al. [25] found that the shortest FAME to exhibit NTC behaviour in the
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 158
C H
3
CH
2
C H
2
C
O C H
3
O
C H
2
C
C H
2
C H
3
CH
C H
C
O C H
3
O
CH
C H
C H
3
CH
C H
2
C H
2
CH
2
C H
3
CH
C H

C H
3
CH
2
C H
2

C H
3
C
C H
Figure 9.4: A comparison of the combustion pathways for methyl trans-2-butenoate
(above) and methyl butanoate (below) which lead to unsaturated hydrocarbons.
JSR was methyl hexanoate. Similarly, Hadjali and coworkers [16] found that methyl
hexanoate exhibited high pressure autoignition delay times comparable to long chain
alkanes (i.e., n-heptane). Therefore, methyl hexanoate was proposed as a suitable sur-
rogate for long chain FAME. Dayma et al. developed a chemical kinetic mechanism for
methyl hexanoate consisting of 435 species and 1875 reversible reactions, and validated it
against data obtained in a JSR [25]. The model was developed by adding reactions to the
Fisher mechanism for MB, and therefore much of the same chemistry with respect to the
ester function was observed. The mechanism indicates that methyl hexanoate is mainly
consumed by H-atom abstraction reactions from the carbon. More interestingly, the
study indicated the consumption of methyl hexanoate proceeds in much the same way
as a straight chain alkane.
In an eort to study the longer chain FAMEs found in real biodiesel, Dagaut and
coworkers [26] studied the oxidation of rape seed oil methyl ester (RME) in a JSR at
various temperatures and pressures. RME is a complex mixture of C
14
-C
18
esters so
developing a chemical kinetic mechanism would be cumbersome. The authors proposed
that a long chain alkane would be a suitable surrogate for RME since experimental data
for n-hexadecane (C
16
H
34
) in the JSR at similar conditions indicated similar product
species concentration proles. Therefore, a detailed chemical kinetic mechanism for n-
hexadecane was used to simulate the RME experiments. This gave a good description
of the RME experimental results, with a good agreement for RME reactivity and the
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 159
relative importance of C
2
-C
6
alkenes. The authors stated that the mechanism could be
improved by including chemical kinetic pathways for the following: i. the ester moiety
to reproduce the early CO
2
formation found in RME, and ii. carbon double bonds to
reproduce large alkene production in RME attributed to unsaturated FAMEs.
9.3.1 Mechanisms for Methyl Decanoate
Methyl decanoate has been proposed as a much better surrogate for biodiesel due to the
length of the alkyl chain. Vaughan et al. [15] found that the ignition time of methyl
decanoate (MD) fuel droplets in microgravity matched well with those of soybean oil
methyl esters. Szybist and coworkers [19] studied the autoignition of MD in a mo-
tored engine and compared it to n-heptane, a petroleum diesel surrogate. The authors
found that MD well reproduced the NTC behaviour that is characteristic of diesel and
biodiesel fuels. MD had a greater heat release during low temperature ignition, which
was attributed to the fully saturated aliphatic chain and not the ester group. Higher
levels of CO
2
were observed at low temperatures, and it was hypothesized that this is
due to decarboxylation of the ester group, similar to pathways observed in kinetic studies
on MB. However, since actual biodiesel may be highly unsaturated, it is likely that using
MD as a surrogate will overpredict the low temperature heat release of actual biodiesel.
Herbinet et al. [27] have developed a detailed chemical kinetic mechanism for MD
consisting of 3012 species and 8820 reactions. Since there are no combustion studies of
MD in fundamental experimental congurations (e.g. shock tubes, laminar ames, stirred
and ow reactors), the mechanism was validated against MD data in motored engine
[19] and rapeseed oil methyl ester oxidation data in a JSR [26]. The MD mechanism
is capable of reproducing the early CO
2
formation observed for RME in the JSR, a
behaviour that the n-hexadecane model by Dagaut et al. [26] could not reproduce.
The chemical kinetic mechanism reveals that low temperature formation of CO
2
comes
directly from the presence of the ester group, and since CO is not formed directly, the
soot reducing eciency of the fuel-bound oxygen is not maximized. The mechanism is
unable to reproduce large alkene production in RME because MD is too small, and it
does not contain the double bonds which lead to alkene formation. In addition, the large
size of this mechanism requires a robust numerical solver and enormous computing power
when attempting to model combustion in some congurations, such as laminar ames.
Zhang and coworkers recently studied the low temperature ignition chemistry of
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 160
three C
9
FAME (i.e., methyl nonanoate, methyl trans-2-nonenoate, and methyl trans-
3-nonenoate), and used Herbinets [27] MD mechanism to interpret their results. It was
revealed that unsaturated FAME are less reactive at low temperatures because the C-
C double bonds inhibit the formation of six- or seven-membered transition state rings,
which are important in low temperature ignition chemistry. The reactivity decreases as
the double bond moves towards the center of the fatty acid chain. The study found that
low temperature ignition chemistry is dependent on the structure of the fatty acid chain
and not on the ester moiety; saturated FAME follow ignition pathways similar to straight
alkanes while unsaturated FAME follow similar pathways as straight alkenes.
The aforementioned MD chemical kinetic mechanism is limited in its applicability
due to the large number of species and reactions. In addition, the chemical stiness,
which is characterized by dramatic dierences between species and reaction time scales,
is signicant for the large molecules in the detailed mechanism [28]. Using this mechanism
in a zero-dimensional simulation (i.e., JSR) is computationally expensive, and henceforth
a one-dimensional simulation (i.e., opposed-ow diusion ame) is virtually impossible.
To overcome these problems, Seshadri and coworkers used the directed relation graph
(DRG) method to reduce the detailed mechanism into a skeletal mechanism consisting
of 713 elementary reactions and 125 species [29]. The skeletal mechanism was capable of
predicting experimental extinction and ignition of MD in an opposed-ow diusion ame.
A large number of low temperature chemical reactions in the original mechanism were
discarded during the reduction, indicating that low temperature chemistry is of minor
importance in an opposed-ow diusion ame.
9.4 Background Summary and Research Motivation
The knowledge of the high and low temperature chemical kinetic reactions responsible
for biodiesel consumption is necessary to simulate ignition, combustion, and emissions
in diesel engines. Developing validated chemical kinetic mechanisms for real biodiesel is
dicult for the following reasons:
Biodiesel is a complex mixture of saturated and unsaturated fatty acid alkyl esters
that varies depending on the feedstock used for its preparation.
Developing mechanisms for the large chain molecules is dicult since the num-
ber of reaction pathways and intermediate species increases drastically with each
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 161
additional carbon atom.
Obtaining the vaporised fuel streams needed in most fundamental combustion se-
tups is challenging for large chain molecules due to their low vapour pressures and
high boiling points.
Therefore, surrogate fuels of well-characterized composition are needed to model
biodiesel combustion. Short chain methyl esters were originally proposed as surrogate
fuels, but detailed experimental and kinetic studies revealed that they do not reproduce
the important low temperature combustion properties of real biodiesel. However, mecha-
nisms for these smaller molecules have deepened our understanding of the ester function
during combustion. Intermediate chain length FAME, and possibly even straight chain
hydrocarbons, appear to be better biodiesel surrogates. Additional experimental and
modeling work is needed to develop validated chemical kinetic mechanisms for saturated
and unsaturated C
8
to C
10
FAME. The chemical kinetic mechanism for these surrogates
would reproduce the distinguishing features of biodiesel, namely the combined eects of
a large carbon chain, carbon double bonds, and the ester moiety.
The goal of this study is to develop an experimentally validated chemical kinetic
mechanism for MD. The existing detailed MD mechanism by Herbinet et al. [27] and
the skeletal mechanism by Seshadri et al. [29] have not been validated against funda-
mental combustion data for MD because experiments have not been performed. This
study presents new experimental temperature and species concentration proles for an
MD opposed-ow diusion ame, and uses this data to develop an improved skeletal
mechanism for MD combustion.
9.5 Experimental Methods
A detailed explanation of the experimental opposed-ow diusion ame and correspond-
ing sampling setup was presented in Chapter 4. Briey, the setup consists of two identical
at ame burners with circular burner ports of diameter 25.4 mm, facing each other and
spaced 20 mm apart. A fuel mixture of 98.2% N
2
and 1.8% fuel (99% pure MD) was
fed through the bottom port at a mass ux rate of 0.0142 g/cm
2
-sec, while an oxidizer
mixture of 42.25% O
2
and 57.75% N
2
was fed through the top port at a mass ux rate of
0.0137 g/cm
2
-sec. At these plug ow conditions, the Reynolds Number is in the laminar
ow regime (i.e. 1c < 400), the ame is on the fuel side of the stagnation plane, and the
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 162
fuel side strain rate is approximately 31 :
1
. An ultrasonic atomizer was used to spray
the liquid fuel into a stream of N
2
gas. The gaseous fuel mixture was delivered to the
burner via heated stainless steel tubing. The temperatures of the gases exiting the top
and bottom burner ports were 420 K and 400K , respectively. The gas sampling system
in these experiments consists of a quartz microprobe (250 m ID, 300 m OD) connected to
a dual-stage pump with heated heads (388 K) containing PTFE diaphragms. The suc-
tion side of the sampling system consisted of 1,4

tubing and a vacuum pressure gauge


connected to the quartz microprobe. An absolute pressure of 4-6 kPa was measured
downstream of the microprobe, and this was sucient to quench most reactions and en-
sure accurate data on ame composition. The compression side of the pump delivered
the samples to the analytical instruments via 1,4

stainless steel tubing heated to 388


K.
Analytical techniques used to measure the species in the sample included: NDIR for
CO and CO
2
; GC/FID with an HP-Al/S PLOT column for C
1
to C
5
hydrocarbons;
and GC/FID equipped with a methanizer (i.e., Ni catalyst) and Poraplot-U column for
oxygenated hydrocarbons such as acetaldehyde/ethenol, formaldehyde, and acrolein. The
precision of species measurements is estimated to be 15%. Temperature measurements
were obtained using a 254 jm diameter wire R-type thermocouple (Pt-Pt/13% Rh) in
an apparatus similar to that used by McEnally et al. [30]. The measured temperatures
were corrected for radiation losses.
9.6 Computational Methods
The kinetic modeling for MD oxidation in the opposed-ow diusion ame was performed
using the OPPDIF code within the CHEMKIN modeling package [31]. The inputs to
each simulation include a detailed chemical kinetic reaction mechanism, a dataset of
thermochemical properties, and a dataset of transport properties.
9.6.1 Chemical Kinetic Mechanism
The chemical kinetic mechanism developed here is an extension of previously published
detailed and skeletal mechanisms for MD. The large size of Herbinet et. als [27] de-
tailed mechanism makes it impractical for use in the one-dimensional ame code (i.e.,
OPPDIF), while the skeletal mechanism proposed by Seshadri et al. [29] does not contain
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 163
enough species and reactions to accurately predict species concentration proles in the
opposed-ow diusion ame. Therefore, the present study develops an intermediate size
mechanism, which balances computational performance and chemical delity. Initially,
several modications were made to the detailed chemical kinetic mechanism to better
represent MD combustion, and then this modied mechanism was reduced using the
DRG method.
Modied Detailed Chemical Kinetic Mechanism
Herbinet et al.s detailed chemical kinetic mechanism includes low temperature chem-
istry to simulate fuel ignition and NTC behaviour, as well as intermediate and high
temperature chemistry to simulate fuel combustion and product species formation. Low
temperature chemistry is not addressed in the present study because the consumption of
fuel in an opposed-ow diusion ame is dominated by high temperature chemical reac-
tions. Therefore, the high temperature part of the detailed methyl decanoate mechanism
and the corresponding modications are discussed here.
For the most part, the high temperature consumption of MD proceeds similarly to a
straight alkane. The decomposition is driven by unimolecular decomposition and H-atom
abstraction reactions leading to alkyl and alkyl-ester radicals. These radicals then react
via isomerization, decomposition (i.e., /ctc-scission) and bimolecular reactions with O
2
.
The unimolecular decomposition reactions were written in the reverse radical-radical
recombination direction and the rate for the decomposition direction was calculated from
thermochemistry via microscopic reversibility. The rates for unimolecular decomposition
reactions were based on a previous mechanism for MB by Fisher et al. [1]. H-atom
abstraction from MD and other hydrocarbon molecules were included for reactions with
radicals (e.g., H, CH
3
, O, OH, etc.), and the rates were determined based on typical
hydrocarbon C-H bond energies for primary, secondary, and tertiary H atoms. The
reaction rates for the two H atoms bonded to the carbon atoms adjacent to the carbonyl
group were based on the mechanism for MB by Fisher et al.
Herbinet et al. performed computational simulations of RME in a JSR using the
detailed MD mechanism and showed that unimolecular decomposition reactions are re-
sponsible for the consumption of fuel at 1040 K [27]. We also conducted opposed-ow
diusion ame simulations using the skeletal mechanism by Seshadri et al. [29] and found
unimolecular decomposition signicant at 1200 K. This predominance of unimolecular de-
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 164
composition is unexpected because chemical kinetic studies on MB [2] and MC [22] in a
JSR and opposed-ow diusion ame and methyl hexanoate [25] in a JSR indicate that
H-atom abstraction reactions are predominant under these conditions. Additionally, ab
initio calculations by Huynh et al. [14] indicate that the energy barrier for unimolecular
decomposition of MB is higher than H-atom abstraction reactions, so H-atom abstrac-
tion reactions would dominate in combustion environments where reactive radicals are
abundant (e.g., ames, premixed reactors, etc.) We ran the JSR simulations again for
the same conditions and found that H abstraction reactions are indeed predominant at
1040 K, so Herbinet et al.s statement that unimolecular decomposition dominates in
the JSR was incorrect. However, the unexpected predominance of unimolecular in the
opposed-ow diusion ame simulations indicated that unimolecular decomposition rates
in original detailed mechanism needed attention.
A recent study on the autoignition of MB by Dooley et. al [4] proposed improved
rates for the unimolecular decomposition of MB into methyl ester plus alkyl radicals.
Additionally, Huynh et al. [14] and Dooley et al. [4] proposed new H-atom abstraction
rates for MB. The present study uses the study by Dooley et al. to develop improved re-
action rates for several MD unimolecular decomposition, H-atom abstraction, and radical
decomposition reactions. It should be noted that using reaction rates determined for MB
can be condently applied to MD because theoretically calculated C-H, C-C, C-O bond
dissociation energies in long chain FAMEs [11, 10] are similar to those calculated for MB
[9]. The following list of modications were made to better represent the combustion of
MD in the opposed-ow diusion ame:
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 165
The detailed mechanism by Herbinet et al. had an error in the activation energy for
H-atom abstractions reactions by OH from secondary C-H bonds. The value was
changed from its erroneous value (i.e., -3500 kcal/mol) to the correct value (i.e.,
-35 kcal/mol) in all relevant reactions. The original authors
1
discovered this error
after publishing the mechanism, but found that it did not alter their results much.
However, the present study shows that the error does have a signicant aect on
the low temperature reactivity of MD.
The recombination rate of 1-octene (C
8
H
16
) and the ME2J radical to form the
MD4J radical (i.e., reverse of the decomposition of the MD4J radical, as shown
Figure 9.5) was changed to
8.80r10
3
1
2.48
exp
(
6130
co|
nc|
11
)
c:
3
:o| :
to make it consistent with rates of analogous reactions for the MB5J, MB6J, MB7J,
etc. radicals.
C
8
C
7
C
6
C
5
C

4
C
3
C
2
1
O
C
O
C
9
C
10
C

O
C
O
C
C
C
C
C
C C
C
+
MD4J
ME2J
C
8
H
16
-1
Figure 9.5: Decomposition of the MD4J radical to 1-octene (C
8
H
16
) and the ME2J radical
The recombination rate of methyl 2-propenoate (MP2D) and the 1-heptyl radical
(C
7
H
15
) to form the MD2J radical (i.e., reverse of the decomposition of the MD2J,
as shown Figure 9.6) was changed to
1.76r10
4
1
2.48
exp
(
8130
co|
nc|
11
)
c:
3
:o| :
based on the rate expression given by Curran et al. [32] for the recombination of
propene (C
3
H
6
) and the methyl radical (CH
3
) to form the 2-butyl radical (sC
4
H
9
)
1
Dr. William Pitz and Dr. Olivier Herbinet, Lawrence Livermore National Laboratory
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 166
(i.e., reverse of the decomposition of the sC
4
H
9
). Currans estimate was modied
by 2 kcal/mol to account for resonance stabilization eects of the carbonyl group
in MD2J [4].
C
8
C
7
C
6
C
5
C
4
C
3
C

2
1
O
C
O
C
9
C
10
C
C O
C
O
MD2J
+
C
7
H
15
CO
Methyl 2-propenoate
(MP2D)
C
C
C
C
C

C
C
Figure 9.6: Decomposition of the MD2J to methyl 2-propenoate (MP2D) and the C
7
H
15
radical
Hydrogen atoms bonded to the alpha carbon (see Figure 9.7) have BDEs similar
to tertiary C-H bonds in alkanes [4]. Therefore, H atom abstraction rates by the
radicals H, OH, CH
3
, CH
3
O, and HO
2
were changed to analagous rates for tertiary
H atom abstraction in isobutane (iC
4
H
10
). The rates for isobutane from Healy et
al. [33] were multiplied by 2 to account for greater number of H atoms in the MD2J
radical.
C
8
C
7
C
6
C
5
C
4
C
3
C
2
1
O
C
O
C
9
C
10
MD
+ R
C
8
C
7
C
6
C
5
C
4
C
3
C

2
1
O
C
O
C
9
C
10
+ RH
MD2J
Figure 9.7: Abstraction of H atoms from the alpha carbon atom by a reactive radical
species (R)
Hydrogen atoms bonded to the methoxy carbon (see Figure 9.8 for MDMJ) have
BDEs similar to secondary C-H bonds in alkanes [4]. Therefore, H atom abstraction
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 167
rates by the radicals H, OH, CH
3
, CH
3
O, and HO
2
from MDMJ amd MP2DMJ
were changed to analagous rates for secondary H atom abstraction in n-butane
(C
4
H
10
). The rates for n-butane from Healy et al. [33] were multiplied by 1.5 to
account for greater number of H atoms in the methyl ester radicals.
C
8
C
7
C
6
C
5
C
4
C
3
C
2
1
O
C
O
C
9
C
10
MD
+ R
+ RH
C
8
C
7
C
6
C
5
C
4
C
3
C
2
1
O
C

O
C
9
C
10
MDMJ
Figure 9.8: Abstraction of H atoms from the methoxy carbon atom by a reactive radical
species (R)
The reaction rates for unimolecular decomposition involving C-C bonds in or around
the carbonyl group (see Figure 9.9) were changed to make them consistent with the
rate of analogous reaction for MB proposed by Dooley et al. [4]. These reactions
were written in the recombination direction in the original MD mechanism, while
Dooley et al. provided the rate constants for MB in the decomposition direction
and treated them for pressure dependency. We rewrote the MD reactions in the
decomposition direction and used Dooley et al.s high pressure rate constant for
MB decomposition. The pressure dependency parameters were removed because
MD is a much larger molecule than MB, so the same pressure dependency would
not apply.
The decomposition of the methoxy radical (CH
3
O) to formaldehyde (CH
2
O) and
the H radical was changed to
1.38r10
21
1
6.65
exp
(
33190
co|
nc|
11
)
c:
3
:o| :
based on the rate proposed by Tsang et al. [34]. This change was made to improve
the prediction of formaldehyde concentrations in the opposed-ow diusion ame.
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 168
C
8
C
7
C
6
C
5
C
4
C
3
C
2
1
O
C
O
C
9
C
10
MD
C

O
C
O
C
8
C
7
C
6
C
5
C
4
C
3
C
2
1
O
C
9
C
10
C
9
H
19
CO
- CH
3
O
- C
9
H
19
CH
3
OCO
C

2
1
O
C
O
- C
8
H
17
ME2J
- C
7
H
15
C
2
1
O
C
O
C
3
MP3J
C
8
C
7
C
6
C
5
C
4
C
3
C
2
1
O

O
C
9
C
10
- CH
3
DAOJ
Figure 9.9: Unimolecular decomposition of MD via scission of C-C bonds in and around
the carbonyl group
Skeletal Chemical Kinetic Mechanism
The modied detailed chemical kinetic mechanism with 3012 species and 8820 reactions
is impractical for use in a one-dimensional ame code. However, mechanism reduction
methods are available to reduce the detailed mechanisms size and complexity. The two
primary methods of reducing a mechanism are: i. elimination of unimportant species and
reactions and/or ii. lumping of species with similar structures (e.g., isomers) and reaction
pathways. When elimination is performed, the resulting mechanism is termed a skeletal
mechanism, and when lumping is performed the result is a reduced mechanism. In this
study, the DRG method is used to generate a skeletal version of the modied detailed
MD mechanism discussed previously.
Seshadri et al. [29] have already used the DRG method to generate an MD skele-
tal mechanism consisting of 125 species and 713 elementary reactions. The researcher
responsible for the development of the skeletal mechanism using the DRG methodology
was Dr. Tianfeng Lu
2
. The reduction of detailed chemical kinetic mechanisms is not
the focus of the present thesis study, so Dr. Lus assistance was sought in developing
2
University of Connecticut, Department of Mechanical Engineering
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 169
a skeletal MD mechanism capable of predicting species and temperature proles in the
opposed-ow diusion ame. Below is a brief description of the DRG method, followed
by a discussion of the skeletal mechanism generated in this study.
The DRG is a systematic algorithm used for eliminating unimportant species and
reactions from a detailed mechanism. In this method a directed graph is used to detect all
the species in the mechanism that are strongly coupled to the fuel and oxidizer [28]. The
directed graph is generated using sampling points covering a wide range of temperatures,
pressures, fuel-oxygen-nitrogen mixture fractions, and uid mixing conditions, such that
the couplings remain valid for combustion in various platforms (i.e., laminar ames, shock
tubes, jet-stirred reactors, etc.). The DRG methodology and its applicability to various
hydrocarbon mechanisms is available in the literature [35, 36].
The rst step in the DRG method is to set up a sample space representing the various
combustion regimes of interest. The detailed mechanism is then run for all the points
in the sample space so that a directed graph of species coupling is generated. From the
directed graph, the user can select the desired error tolerance (e.g., 10%-30%), such that
unimportance species can be eliminated.
The previously proposed skeletal mechanism was capable of predicting experimental
extinction and ignition of MD in an opposed-ow diusion ame with reasonably good
accuracy. However, its ability of predicting species proles in an JSR and an opposed-ow
diusion ame is poor. Figure 9.10 displays JSR simulations using various FAME and
alkane mechanisms at o=1.0, P=1013.25 kPa, t=1 s, 0.1% fuel mole fraction. n-Decane
simulations [37] and experimental data [38] match well, with both showing cool ame
reactivity in the range of 600-800 K. Simulations using Dooleys MB mechanism [4] and
Seshadri et al.s skeletal MD mechanism [29] indicate that these models lack cool ame
behaviour, and therefore they are not suitable for modeling real FAME low temperature
chemistry.
It is observed that Herbinet et al.s detailed MD mechanism [27] displays cool ame
reactivity, but when compared to simulations for n-decane [37], the MD model appears
to overpredict the fuels reactivity. One would expect the reactivity of MD and n-decane
in the JSR to be similar since both contain C
10
alkyl chains, and shock tube studies
also indicate that their reactivity is similar [27]. This study identied the problem as
incorrect activation energies for H-atom abstractions reactions by OH from secondary
C-H bonds, as mentioned previously. After correcting these values, the modied detailed
MD mechanism well predicts the reactivity of n-decane.
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 170
1.E-05
1.E-04
1.E-03
1.E-02
500 600 700 800 900 1000
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
MB (Dooley et al., 2008) MD (Seshadri et al., 2009)
MD (Herbinet et al., 2008) n-decane (Dagaut et al., 1995)
n-decane (Westbrook et al., 2008) MD (Sarathy et al., detailed)
MD (Sarathy et al., skelatal)
Figure 9.10: Experimental (symbols) and computed (lines with symbols) proles obtained
from the oxidation of methyl decanoate, n-decane, and methyl butanoate in a JSR at
o=1.0, P=1013.25 kPa, t=1 s, 0.1% fuel mole fraction.
Figure 9.11 compares the skeletal MD mechanism of Seshadri et al., the detailed MD
mechanism by Herbinet et al. [27], and experimental data for RME in a JSR [26]. The
125 species mechanism by Seshadri et al. poorly predicts the concentrations of carbon
dioxide, carbon monoxide, and methane, especially at lower temperatures. This indicates
that the mechanism lacks chemical delity when compared to the detailed mechanism.
After discussion with Dr. Lu, it was decided that the sample space be modied to
produce an improved MD skeletal mechanism. The chosen sample space included the
following:
Combustion in a closed homogeneous gas phase reactor at pressures of 101.325 kPa
and 1013.25 kPa, temperatures spanning 900-1800 K, equivalence ratios spanning
o=0.25-2.0, with mixtures of undiluted fuel plus air, and diluted fuel (i.e. 2% MD,
98% N
2
) plus air. These conditions were simulated using the SENKIN code in
CHEMKIN.
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 171
Combustion in an open perfectly stirred reactor at pressures of 101.325 kPa and
1013.25 kPa, temperatures spanning 900-1500 K, equivalence spanning o=0.25-2.0,
residence times of 0.0001 s, 0.001 s , 0.01 s, and 1 s, and mixtures of undiluted fuel
plus air. These conditions were simulated using the PSR code in CHEMKIN.
The directed graph generated from the sampling points at the aforementioned con-
ditions was used to generate a skeletal mechanism consisting of 648 species and 2998
reactions. This skeletal mechanism accurately reproduces the low temperature reactiv-
ity predicted by the detailed mechanism, as shown in Figure 9.10. In addition, species
proles for RME in the JSR, as shown in Figure 9.11, are also very similar for this
new skeletal mechanism and the detailed mechanism. This indicates that the proposed
skeletal mechanism is acceptable for modeling detailed chemical kinetic processes in the
JSR. Furthermore, it is suitable replacement to the detailed mechanism for predicting
the combustion properties of MD.
1.E-05
1.E-04
1.E-03
1.E-02
790 890 990 1090 1190
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
carbon monoxide
carbon dioxide
methane
ethylene
Herbinet et al., 2008
Seshadri et al, 2009
Sarathy et al, this study
Figure 9.11: Comparison of MD mechanisms (lines with symbols) and experimental data
(symbols) for RME in a JSR at o=1.0, P=101.325 kPa, t=1.0 s [26].
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 172
9.6.2 Thermochemical Data
The thermochemical data for MD published by Herbinet et al. [27] was used in this study.
The thermochemical properties for molecules and radicals were calculated using THERM
[39], which is a software based on the group and bond additivity methods proposed by
Benson [40]. THERM determines the thermochemical properties of a radical species by
applying a bond dissociation (BD) group increment to a stable molecule which reects the
loss of an H atom from that species. The BD groups are based on specic bond energies
and dierence in heat capacities and entropies for specic molecular classes [39]. For
more accurate calculations, the user can modify BD group values for specic molecular
classes based on ab initio thermochemical calculations, such as those by Sumathi and
Green [41].
With the assitance of Dr. Bill Pitz, a coauthor of the Herbinet et al. MD mechanism
[27], we found that the BD groups used in the calculation of thermochemical properties for
methyl ethanoate (ME), and it radicals ME2J and MEMJ, were incorrect in the original
work. In this study, the thermochemical properties were updated based on ME2J bond
energies calculated by El-Nahas et al. [9]. The new thermochemical parameters make
these molecules more stable and decrease their decomposition rates.
9.6.3 Transport Properties
The molecular transport parameters for species were obtained using a variety of methods.
This study builds upon the transport property database developed by Seshadri et al.
[29] for the 125 species contained within their MD skeletal mechanism. The transport
properties of species with no previously published data were determined as follows.
Most of the transport parameters that needed to be determined were for stable C
2
-C
10
saturated and unsaturated methyl esters and their corresponding radicals. We assume
that the transport properties are similar for saturated and unsaturated methyl esters of
the same chain length, so we only performed calculations for saturated methyl esters and
used the same values for their unsaturated counterparts. For methyl ester radical species,
the transport properties of their stable counterpart were used.
For the stable saturated methyl ester species, this study used the correlations de-
veloped by Tee, Gotoh, and Stewart [42], as described by Wang and Frenklach [43], to
calculate the Lennard-Jones collision diameter and potential well depth using the P
c
,
T
c
, and T
o
of the species. For C
2
-C
6
methyl esters, these values were obtained from
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 173
the NIST Chemistry WebBook [44], but the P
c
, T
c
values were not available for C
7
-C
10
methyl esters. However, a strong correlation exists between carbon chain length and P
c
and T
c
for C
3
-C
6
methyl esters, as shown in Figure 9.12, so we extrapolated the P
c
and
T
c
for larger methyl esters using a power law function.
100
200
300
400
500
600
700
0 2 4 6 8 10 12
Carbon Number
C
r
i
t
i
c
a
l

T
e
m
p
e
r
a
t
u
r
e

(
K
)
20
25
30
35
40
45
50
C
r
i
t
i
c
a
l

P
r
e
s
s
u
r
e

(
P
a
)
Measured Tc
Extrapolated Tc
Measured Pc
Extrapolated Pc
Figure 9.12: Critical pressure (P
c
) and critical temperature (T
c
) for C
2
-C
10
methyl esters.
Experimentally measured dipole moments were obtained from Mcclellans text [45].
This property was available for C
3
-C
6
methyl esters but not for C
7
-C
10
methyl esters.
The molecular dipole moment is a vector property that can be determined for an un-
known molecule using vector addition of known bond moments [46]. However, such a
method requires detailed information of the geometry of molecular bonds and their elec-
tronegativities (i.e., polarity). The problem is simplied in the present study because
experimentally measured dipole moments for saturated C
3
-C
6
methyl esters fall in the
range of 1.61-1.76 Debyes. This indicates that the molecular dipole moment is created
by the ester moiety, which exhibits a strong polarity, and not the saturated alkyl chain,
which is nonpolar. Therefore, a dipole moment of 1.70 Debyes was used for the C
7
-C
10
methyl esters.
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 174
Experimentally measured values for polarizability (c) in cubic Angstroms (

A
3
) were
obtained from the CRC Handbook of Chemistry and Physics [47]. The polarizability
can also be determined using the empirical relation proposed by Bosque and Sales [48],
which allows estimation using the molecular formula (i.e., # of C, H, and O atoms), as
shown in Equation 9.1. Table 9.1 presents experimentally and empirically determined
polarizabilities for several FAME. The calculated values are within 1% of the measured
values for C
3
-C
6
methyl esters, so a high degree of condence accompanies the calculated
polarizabilities for C
7
-C
10
methyl esters.
c = 0.32 + 1.51 #C + 0.17 #H + 0.51 #C (9.1)
Table 9.1: Experimentally and Empirically Determined Polarizabilities (

A
3
) for FAME
Molecular Formula Experimental [47] Empirical [48]
Methyl ethanoate C
3
H
6
O
2
6.94 6.89
Methyl propanoate C
4
H
8
O
2
8.97 8.74
Methyl butanoate C
5
H
10
O
2
10.41 10.59
Methyl pentanoate C
6
H
12
O
2
- 12.44
Methyl hexanoate C
6
H
14
O
2
- 14.29
Methyl heptanoate C
7
H
16
O
2
- 16.14
Methyl octanoate C
8
H
18
O
2
- 17.99
Methyl nonanoate C
9
H
20
O
2
- 19.84
Methyl decanoate C
10
H
22
O
2
- 21.69
9.7 Results and Discussion
9.7.1 Opposed-ow Diusion Flame
The proposed skeletal MD mechanism was validated against experimental data obtained
in an MD opposed-ow diusion ame. The opposed-ow diusion ame allows the
study of fuel oxidation in a non-premixed laminar ame environment. Concentration
proles for species were obtained by sampling the product gas at various points between
the two burner ports and then analyzing it using a variety of analytical techniques.
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 175
The measured species included methyl decanoate (MD), carbon monoxide (CO), car-
bon dioxide (CO
2
), formaldehyde (CH
2
O), methane (CH
4
), acetylene (C
2
H
2
), ethylene
(C
2
H
4
), ethane (C
2
H
6
), ketene (CH
2
CO), propane (C
3
H
8
), propene (C
3
H
6
), propyne
(pC
3
H
4
), 1-butene (1-C
4
H
8
), 1,3-butadiene (1,3-C
4
H
6
), 1-pentene (1-C
5
H
10
), 1-hexene
(1-C
6
H
12
), 1-heptene (C
7
H
14
), and 1-octene (1-C
8
H
16
). A species prole was identied
for C
2
H
4
O, but we are unable to determine if the compound is ethanal (i.e., acetaldehyde)
(CH
3
CHO) or ethenol (C
2
H
3
OH) since both have the same retention time on the GC
column. In addition, ethenol rapidly tautomerizes to acetaldehyde upon contact with
surfaces [49, 50], so we assume that the C
2
H
4
O measured in the GC is the combined
concentration of acetaldehyde and ethenol in the ame. Species below the experimental
limit of detection (LOD) (i.e., 5 ppm) included 1-butyne, n-butane, 2-butyne, trans-2-
butene, cis-2-butene, pentane, hexane, propanal, 2-propenal (i.e., acrolein), 2-propanone
(i.e., acetone), butanal, methyl 2-propenoate, methyl 3-butenoate, methyl 4-pentenoate,
and methyl 5-hexenoate.
9.7.2 Temperature, Fuel, and Hydrocarbon Species
Figure 9.13 displays the measured and predicted species and temperature proles ob-
tained in the opposed-ow diusion ame. The experimental results (solid symbols)
show that the MD concentration begins decreasing quickly at a distance of 5 mm from
the fuel port. As the fuel is consumed, the CO and CO
2
concentrations begin rising.
All of the MD is consumed at a distance of approximately 8.25 mm from the fuel port,
which corresponds closely the visually observed ame front. Both the temperature and
CO
2
concentrations reach their maximum at approximately 9.5 mm from the fuel port.
Just before the ame front, at around 7.75 mm from the fuel port, the concentrations
of hydrocarbon species reach their maximum. Besides CO and CO
2
, the most abundant
measured species are C
2
H
4
, C
2
H
2
, CH
4
, and C
3
H
6
.
It should be noted that these results do not display the shift in the measured species
proles towards from the fuel port, which was observed previously in n-butanol experi-
ments (refer to Chapter 7). This shift was caused by a poor probe design which induced
ow eld disturbances. The results for MD used a new probe design with did not disturb
the ame
3
.
3
The dierent probe designs are described in Chapter 4
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 176
This probe eect was resolved by moving to a new probe design
4
.
The models prediction of temperature species proles in the opposed-ow diusion
is also shown in Figure 9.13
5
. The model reproduces the experimentally measured
temperature prole very well. The reactivity of MD is also well predicted by the model.
The maximum concentration of CO
2
is underpredicted by approximately 0.3%, while the
maximum concentration of CO is underpredicted by approximately 0.1%.
The model performs well qualitatively, in that it well reproduces the shape of the
experimental proles and the points of maximum measured concentrations. In the fol-
lowing discussion, the models prediction is considered good if predicted maximum mole
fraction is within a factor 1.5 of the measured maximum mole fraction. The model per-
forms well at predicting the maximum concentrations of CH
4
, C
2
H
6
, C
3
H
4
, C
3
H
6
, 1-C
4
H
8
,
C
8
H
16
, C
5
H
10
, C
6
H
12
, C
7
H
14
, and 1,3-C
4
H
6
. The model moderately underpredicts (i.e.,
1.5-2 times) the maximum concentration of C
2
H
4
and overpredicts the concentration of
C
2
H
2
. Both model and the experimental data indicate that the concentration of 1-alkenes
decreases with increasing carbon number (e.g., [C
2
H
4
][C
3
H
6
][1-C
4
H
8
][C
5
H
10
] etc.).
A reaction path analysis was performed for MD at 1040 K, the temperature at which
approximately where 50% of fuel is consumed. Approximately 98% of the fuel is consumed
via H-atom abstraction by H atoms (57%), OH radicals (5%), and CH
3
radicals (30%).
Abstraction is favoured for H atoms bonded to the methoxy carbon (20%) and the c
carbon (16%), and then secondary (8% each) and tertiary H atoms (3%).
Figure 9.14 shows the fate of the MDMJ radical which is formed via H-atom abstrac-
tion from the methoxy carbon. Interestingly, -scission is not favored for this radical
because it requires the breaking of a strong C-O bond. Instead, the MDMJ radical un-
dergoes isomerization to form the MD2J radical (61%) and the MD3J radical (36%).
The fate of the MD2J radical is shown in Figure 9.15, while that of the MD3J radical is
discussed later.
As shown in Figure 9.15, H-atom abstraction from the c carbon leads to the forma-
tion of the MD2J radical, which undergoes -scission (99%) to form a 1-heptyl radical
and methyl 2-propenoate. The 1-heptyl radical eventually leads to the formation of
ethylene, 1-pentene, 1-butene, and the radicals C
3
H
7
and C
2
H
5
, while the fate of methyl
2-propenoate is discussed in the next section. The measureable levels of ethylene, 1-
4
The new probe design is described in Chapter 4 and improved results for methyl decanoate opposed-
ow diusion ames are shown in Chapter 9
5
Appendix C contains larger versions of the graphs shown here
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 177

0
200
400
600
800
1000
1200
1400
1600
1800
2000
0 2 4 6 8 10 12 14 16 18 20
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
measured
corrected
predicted

0%
2%
4%
6%
8%
10%
0 2 4 6 8 10 12 14 16 18 20
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
%
)
CO2
CO
MD

0
2000
4000
6000
8000
10000
12000
14000
0 2 4 6 8 10
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
) C2H4
C2H2
CH4

0
200
400
600
800
1000
1200
1400
1600
1800
2000
0 2 4 6 8 10
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C2H6
1-C4H8
C3H6
C8H16

0
50
100
150
200
250
300
350
0 2 4 6 8 10
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C3H4
C5H10
C6H12
C7H14
1,3-C4H6

0
100
200
300
400
500
600
700
800
900
1000
0 2 4 6 8 10
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C2H4O
CH2O
CH2CO

Figure 9.13: Experimental and computed proles obtained from the oxidation of MD in
an atmospheric opposed-ow ame (1.8% MD, 42% O
2
).
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 178
MDMJ
C
8
C
7
C
6
C
5
C
4
C
3
C
2
1
O
C

O
C
9
C
10
- 3%
+ C
9
H
19
CO
O
CH
2
MD2J
- 58%
- 39%
C
8
C
7
C
6
C
5
C
4
C

3
C
2
1
O
C
O
C
9
C
10
MD3J
.....
C
3
C

2
1
O
C
O
Figure 9.14: Reaction pathway diagram for consumption of the MDMJ radical in the
opposed-ow diusion ame at T=1040 K.
pentene, and 1-butene in the ame experiments agree with this reaction path analysis.
Approximately 56% of the fuel is consumed via abstraction of secondary H atoms from
the #4 to #9 carbon atoms. An example of the subsequent reaction pathways is shown
in Figure 9.16 for the MD4J radical. Approximately, 56% of the radical decomposes to
form 1-octene and the ME2J radical, while another 44% leads to the 1-pentyl radical
and methyl 4-pentenoate. The fate of the 1-pentyl radical is displayed in Figure 9.15,
which identies ethylene as the nal product species. The model indicates that methyl
4-pentenoate is consumed via unmimolecular decomposition to form an allyl radical (i.e.,
C
3
H
5
) and the ME2J radical. The ME2J either isomerizes (90%) to form the MEMJ
radical or decomposes (10%) to form ketene and a formaldehyde (i.e., via the methoxy
radical). The MEMJ radical goes to form formaldehyde and carbon monoxide (i.e., via
the acetyl radical). In this way, most of the MD4J radical goes to form formaldehyde,
carbon monoxide, ethylene, 1-octene, and ketene, all of which were measured in the
opposed-ow diusion ame.
Other radicals formed via abstraction of secondary H atoms from other carbons in the
alkyl chain follow a similar path as MD4J. The radicals decompose via two routes, one
leading to an alkene and a methyl ester radical, and the other forming an unsaturated
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 179
MD2J
C
8
C
7
C
6
C
5
C
4
C
3
C

2
1
O
C
O
C
9
C
10
- 99%
C
C O
C
O
C
C
C
C
C

C
C
+
MP2D
- 17%
- 38%
C
C
C
C
C

C H
2
CH
2
C
C

C
C
C
C
C
- 100% - C
2
H
5
C
C
C
C C
- 77%
- C
2
H
4
1-C
5
H
10
C
C
C

C
C
C
C
- C
3
H
7
- 44%
- 85%
C
C
C
C
1-C
4
H
8
- C
3
H
7
C
2
H
4
Figure 9.15: Reaction pathway diagram for consumption of the MD2J radical in the
opposed-ow diusion ame at T=1040 K.
methyl ester and an alkyl radical. The alkyl radicals eventually result in the formation of
1-alkenes, and as shown previously the concentration of 1-alkenes decreases with increas-
ing carbon number. The model predicts that the unsaturated methyl esters are consumed
mainly by unimolecular decomposition to form an allyl radical and a saturated methyl
ester radical that is three carbon atoms shorter (e.g., methyl 6-heptenoate decomposes
to the radical methyl

4-butanoate, methyl 7-octenoate decomposes to the radical methyl

4-pentanoate, etc.). These methyl ester radicals with the radical site on the terminal car-
bon undergo -scission to form ethylene and smaller methyl ester radicals. The process
continues until the radical site nears the carbonyl group and the radical decomposes to
a low molecular weight oxygenated species.
9.7.3 Oxygenated Species
Table 9.2 presents the maximum predicted and measured mole fractions of several unsat-
urated methyl esters, aldehydes, enals, ketones, ketenes, and enols. The measured and
predicted concentrations of oxygenated product species can add insight into the role of
the ester moiety during combustion. The following is a discussion of several important
oxygenated species and their chemistry in the ame.
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 180
MD4J
C
8
C
7
C
6
C
5
C

4
C
3
C
2
1
O
C
O
C
9
C
10
- 90%
C

O
C
O
ME2J
- 44%
- C
5
H
11
C
C
C
C O
C
O
MF4D
- 73%
- C
3
H
5
- 56%
C O
C

O
MEMJ
- 10%
O

CH
3
C H
2
C
O
+
- 100%
O
CH
2
+
C
C

O
- 98%
- 99%
- H
- CH
3
O
C
C
C
C
C
C
C C
C
+
1-C
8
H
16
Figure 9.16: Reaction pathway diagram for consumption of the MD4J radical in the
opposed-ow diusion ame at T=1040 K.
The model performs well at predicting the maximum concentrations of CH
2
CO, but
overpredicts the maximum concentrations of formaldehyde (CH
2
O) and underpredicts
acetaldehyde + ethenol (C
2
H
4
O). The measured concentration of CH
2
CO (i.e., ketene)
was 413 ppm, and the predicted concentration is 315 ppm. This is the rst time ketene
concentrations have been measured in combustion studies of FAME. Figure 9.16 displays
the primary pathway which forms 83% of the ketene in the ame at 1040 K; the methyl
ester radical, ME2J, undergoes -scission to form ketene and methoxy radical (CH
3
O).
Therefore, it is observed that the ester moiety contributes to the formation of ketene.
Approximately 86% of formaldehyde is formed via the decomposition of various
methyl ester radicals with a radical on the methoxy site, such as MDMJ, MEMJ, and
MP2DMJ. As shown in Figure 9.16 for MEMJ, these fuel radicals undergo -scission to
form formaldehyde. Therefore, it is observed that the ester moiety contributes to the
formation of formaldehyde.
The maximum predicted concentration of formaldehyde is more than 2 times greater
than the measured concentration. This discrepancy can be attributed to either exper-
imental errors or modeling inaccuracies, so both are discussed here. The experimental
measurements for formaldehyde were performed using a GC/FID equipped with a meth-
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 181
Table 9.2: maximum Measured and Predicted Concentration of Oxygenated Species
Species Name Measured (ppm) Predicted (ppm)
in Mechanism
Formaldehyde CH2O 319 924
Ketene CH2CO 413 319
Ethanal+Ethenol CH3CHO+C2H3OH 100 15
Propanal C2H5CHO <LOD <1
2-propenal C2H3CHCO <LOD 44
2-propanone C3H6O <LOD 9
Methyl 2-propenoate MP2D <LOD 1149
Methyl 3-butenoate MB3D <LOD 120
Methyl 4-pentenoate MF4D <LOD 37
Methyl 5-hexenoate MH5D <LOD 29
anizer, which break the aldehydes C-O bond and replaces it with a C-H bond. This
method for detecting formaldehyde has yielded good results in JSR studies of FAME
[2, 22, 25, 51]. However, extractive sampling measurements in ames [2, 22, 52, 53] have
yielded similar discrepancies between measured and predicted formaldehyde, and one
group [53] suggests that formaldehyde may be lost due to polymerization in the sampling
lines.
Modeling overpredictions of formaldehyde concentrations have also been observed in
opposed-ow diusion ames of MB [2, 4]. For MD at 1040 K, the present model predicts
that 86% of the formaldehyde is formed via decomposition of the metyl ester radical with
a radical on the methoxy site. A sensitivity analysis on formaldehyde at 1040 K revealed
that the predicted formaldehyde concentration is sensitive to the decomposition rates of
MDMJ, MEMJ, and MP2DMJ. The current rate estimate for the decomposition of these
radicals to formaldehyde are rough estimates, so detailed studies may reveal better rate
constants. Furthermore, the present model indicates that approximately 20% of CO in
the ame is formed directly from formaldehyde via the HCO radical. Since CO concen-
trations are underpredicted by approximately 1000 ppm, the 500 ppm overprediction of
formaldehyde by the model may be corrected by improving rate constants for reactions
linking CO and CH
2
O. Such modications are beyond the scope of the present study,
and would require further investigation into fundamental experimental and theoretical
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 182
rate studies.
Both the model and experiments indicate that the C
3
oxygenated species propanal,
2-propenal (i.e. acrolein), and 2-propanone (i.e. acetone) are not important product
species during the combustion of MD. However, a number of compounds are predicted
at appreciable levels in the ame, but were not detected in the experiments. It should
be noted that the ame sampling and analytical methods were capable of detecting
and quantifying these compounds. It is unlikely that these compounds were lost in the
sampling system since condensation and reactions in the sampling lines were minimized
by operating at low sampling pressures and using transfer lines heated to 200

C.
The major discrepancy is for unsaturated methyl ester species. The experiments
did not measure detectable levels of any unsaturated methyl esters despite having the
appropriate sampling and analytical methods in place to measure them. Oxygenated
hydrocarbons typically create lower signal responses on an ame ionization detector, but
typically this aects detection levels of low molecular weight oxygenates (e.g., formalde-
hyde, acetaldehyde, etc.). Microliter injections of these unsaturated FAME veried that
the analytical instrument used in this study was suitable for their detection. Furthermore,
the accurate measurement of methyl decanoate in the ame samples indicates that the
sampling apparatus was adequate for high molecular weight FAME. It should be noted
that unsaturated methyl esters have been measured in experimental studies of methyl
hexanoate in a JSR at 10 atm[25], albeit at low concentrations.
Methyl 2-propenoate (i.e., MP2D) is the unsaturated FAME predicted in the high-
est concentration. Figure 9.17 displays the production and consumption pathways for
MP2D in the opposed-ow diusion ame at 1040 K. 94% of the MP2D is formed via
-scission of various methyl ester radicals with a radical site on the c carbon (e.g., MD2J,
etc.). The rest is formed via decomposition of the MP3J radical. Methyl 2-propenoate is
then consumed via H-atom abstraction reactions leading to the methyl ester radical with
a radical on the methoxy carbon (i.e., MP2DMJ). H abstraction is favoured from the
methoxy site because the H atoms bonded to the vinylic carbons have higher BDEs. The
MP2DMJ radical undergoes -scission to ultimately form acetylene, carbon monoxide,
formaldehyde. High concentrations of methyl 2-propenoate are predicted because multi-
ple pathways lead to various methyl ester radicals with a radical site on the c carbon,
and these form methyl 2-propenoate faster than it can be consumed via H-atom abstrac-
tion reactions. The rate parameters for the methyl 2-propenoate consumption have been
determined based on analogies with saturated methyl ester molecules and unsaturated
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 183
hydrocarbons. Fundamental experimental and theoretical studies on H abstraction and
unimolecular decomposition reactions may improve the predicted concentration of methyl
2-propenoate.
Another explanation for the discrepancy between the model and predicted values is
possible decomposition of methyl 2-propenoate upon contact with hot surfaces in the
sampling line. Such a mechanism is plausible given the unstable nature of unsaturated
FAME, especially those of low molecular weight. It is possible that FAME are reacting
in the sampling line to form acetaldehyde+ethenol compounds which were measured
in appreciable quantities, but not predicted to be signicant by the model. Further
investigation of the reactivity of methyl 2-propenoate at the conditions in the sampling
line (i.e., T=388 K, P=6-8 kPa) is required.
MP2D
C
C O
C
O
.....
C
3
C

2
1
O
C
O
- CH
2
O
+ 94%
C
C O
C

O
- H - 90 %
- 100 %
C
C
C

O
MP2DMJ
- C
x
H
y
- 100 %
- C
2
H
3
O
C
MX2J
Figure 9.17: Reaction pathways for the formation and consumption of methyl 2-
propenoate in the opposed-ow diusion ame at T=1040 K.
The Fate of the Ester Moiety
Previous studies on methyl butanoate combustion discussed the fate of the ester moiety
at high temperatures. When applied to dierent experimental conditions (e.g., ames,
premixed reactors, etc.), the MB models predict that the methoxycarbonyl radical is an
important combustion intermediate. We ran the opposed-ow diusion ame simulations
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 184
for MB using the experimental conditions of Gail et al. [2] and the mechanism of Dooley
et al. [4] to calculate how much of the fuel decomposes to form the methoxycarbonyl
radical at high temperatures. The results indicate that approximately 23% of the fuel
ends up in the methoxycarbonyl radical, as shown in Figure 9.18. In Figure 9.3, it
was shown that the methoxycarbonyl radical primarily decays to form a methyl radical
and CO
2
. From a soot reduction standpoint this decarboxylation of the ester is not an
ecient use of fuel-bound oxygen because two oxygen atoms are bonded to one carbon
atom. It would be more ecient if the ester moiety led to the production of CO since each
oxygen atom in the ester group would sequester one carbon atom from participating in
the production of soot. Many researchers have hypothesized that the long chain FAME in
biodiesel undergo similar reaction pathways as methyl butanoate, and therefore the ester
moiety in biodiesel is not as ecient at supressing soot compared to other oxygenated
moieties (e.g., aldehydes, ethers, alcohols, etc.). The present study on methyl decanoate
oers additional insights to the aforementioned hypothesis.
MB
C 4
C
3
C
2
1
O
C
O
- 27%
- H
- 16%
- H
C
C

C O
C
O
MB3J
- 84%
- C
3
H
6
C

O
C
O
CH
3
OCO
C 4
C
3
C

2
1
O
C
O
C
C O
C
O
MB2J
MP2D
- 84%
sum of all
pathways
- 40%
- CH
3
Figure 9.18: Reaction pathways leading to the formation of the methoxycarbonyl radical
in the opposed-ow diusion ame at T=1030 K given the experimental and modeling
conditions of Gail et al. [2].
The above analysis on MB indicates that the radical sites on the c and carbons can
lead to the methoxycarbonyl radical. At 1040 K, approximately 8% of MD is consumed
via H-atom abstraction from the carbon leading to the MD3J radical. In addition, 8%
of MD leads to the MD3J radical via the MDMJ radical, as shown in Figure 9.14 . This
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 185
is 30% less than what was observed for methyl butanoate [2] under similar conditions, or
even for methyl hexanoate in a JSR at 950 K [25]. Similarly, 16% of MD is consumed via
H-atom abstraction from the c carbon whereas it is 27% for MB. These pathways in MD,
and for longer chain FAME too, becomes less important because the number of H-atom
abstraction sites increases with chain length. Furthermore, after absrtraction from the
carbon, the resulting MD3J radical undergoes -scission at equal rates to form either i.
1-nonene plus the methoxycarbonyl radical or ii. methyl 3-butenoate plus and a 1-hexyl
radical. However, the analogous radical in MB (i.e., MB3J) strongly favours the route
leading to the methoxycarbonyl radical, since -scission forming methyl 3-butenoate is
thermodynamically unfavourable for the MB3J radical. The importance of fuel radical
isomerization reactions are become more important as the size of the FAME increases.
Similar to the previous studies on MB, the present MD mechanism predicts that the
methoxycarbonyl radical decomposes to form CO
2
. It is estimated that approximately
6% of the fuel ends up forming the CO
2
via the methoxycarbonyl radical, which is much
lower than previous estimates based on MB (i.e., 23%). This number is expected to be
even lower in longer chain FAME because of the increased number of potential pathways
of fuel consumption.
It should be noted that the amount of fuel that ends up in the methoxycarbonyl
radical depends on the experimental conditions (i.e., temperature, pressure, mixture
fraction, etc.). The above analysis applies to high temperature oxidation where H-atom
abstraction is predominant. Simulations should be conducted for conditions in which uni-
molecular decomposition is predominant for additional insights. In any case, this study
suggests that decarboxylation of the ester group in long chain FAME is not signicant.
In fact, much of the original fuel bound oxygen leads directly to oxygenated species such
as carbon monoxide, formaldehyde, and ketene, wherein one oxygen atom is bonded to
one carbon atom. Therefore, the soot reducing eciency of long chain FAME may be
better than previously believed.
9.7.4 Jet Stirred Reactor
The proposed skeletal mechanism for MD was also validated against experimental JSR
data for RME at o=1.0, P=101.325 kPa, t=1.0 s [26]. Since MD is a smaller molecule
than RME, the inlet mole fraction of MD was proportionally increased to match the inlet
carbon ux of RME, as described by Herbinet et al. [27]. The comparison between the
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 186
model predictions and experimental data is shown in Figures 9.19 and 9.20.
The skeletal mechanism performs similarly to the detailed MD mechanism proposed
by Herbinet et al. [27]. The concentrations of CO
2
, CO, C
2
H
4
, O
2
, H
2
, CH
4
, and C
3
H
6
are well predicted by the proposed skeletal mechanism. However, the concentrations of
1-C
4
H
8
, 1-C
5
H
10
, and 1-C
6
H
12
are overpredicted by the model, which was also observed
by Herbinet et al. [27]. The prediction of alkenes is incorrect because RME consists of
longer chain FAME with various degrees of unsaturation, while MD is fully saturated
and has a small chain. Although there is no data available, RME is likely to have larger
carbon chains leading to the formation of larger 1-alkenes (e.g., C
8
) than would MD;
therefore, MD over predicts the concentrations of the smaller 1-alkenes (e.g., C
4
-C
6
).
1.E-05
1.E-04
1.E-03
1.E-02
790 890 990 1090 1190
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
O2 CO
CO2 H2
CH4 C2H2
co2 detailed o2 detailed
Figure 9.19: Comparison of proposed MD skeletal mechanism and experimental data for
RME in a JSR at o=1.0, P=101.325 kPa, t=1.0 s [26].
9.8 Conclusions and Recommendations
This study is the rst to present experimental data for methyl decanoate combustion
that can be used for validating chemical kinetic mechanisms. The combustion of MD
in the opposed-ow diusion ame generates typical hydrocarbon combustion products
(e.g., CO, CO
2
, CH
4
, C
2
H
4
, etc.). Of particular interest is the production of C
5
-C
8
1-
alkenes which are formed after -scission of fuel radicals. The production of low molecular
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 187
1.E-07
1.E-06
1.E-05
1.E-04
1.E-03
790 890 990 1090 1190
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C3H6
1-C4H8
1-C5H10
1-C6H12
Figure 9.20: Comparison of proposed MD skeletal mechanism (lines with symbols) and
experimental data (symbols) for RME in a JSR at o=1.0, P=101.325 kPa, t=1.0 s [26].
weight oxygenated compounds such as formaldehyde, ketene, and isomers of C
2
H
4
O is
also observed. The absence of acrolein and acetone in the measured data suggests that
FAME combustion does not produce these species.
The experimental data presented herein was used to validate an improved skeletal
mechanism for the high temperature oxidation of MD. Initially, modications were made
to a previously proposed detailed mechanism for MD, and then an improved DRG al-
gorithm was performed to create a skeletal mechanism. This new skeletal mechanism
provides excellent qualitative prediction of experimentally measured species and temper-
ature proles in the MD opposed-ow diusion ame. This study highlights the eective-
ness of the DRG method in producing a mechanism that is computationally practical for
one-dimensional ame simulations yet also retains a high level of chemical delity. The
only major discrepancy between the proposed mechanism and the experimental data was
for methyl 2-propenoate, and it is suggested that the experimental sampling apparatus
be reassessed for accurate measurement of this compound.
The validated MD mechanism provides many new insights into the combustion chem-
istry of FAME. Firstly, the combustion of long chain FAME proceeds in much the same
way as a long chain alkane; however, the ester moiety does introduce unique combustion
pathways. It was shown that MD displays dierent combustion pathways than smaller
Chapter 9. Chemical Kinetic Modeling of Biodiesel Combustion 188
methyl esters, such as MB, due to the greater number of possible reaction pathways. This
study suggests that fuel bound oxygen in MD (i.e., the ester moiety) leads primarily to
the formation of carbon monoxide and low molecular weight oxygenated compounds, and
very little actually ends up in the form of CO
2
. Therefore, under the conditions of this
experimental and modeling study, the oxygen in the ester moiety does a good job of
sequestering carbon from participating in soot production. However, these ndings need
to be veried across other fundamental combustion platforms with varying temperatures,
pressures, mixture fractions, and mixing conditions.
The proposed mechanism also indicates that unsaturated methyl esters are an im-
portant intermediate in the combustion of saturated FAME. The present mechanism
derives kinetic information for unsaturated methyl esters from analogous reaction path-
ways for alkenes. This approximation needs to be improved via fundamental studies on
the thermochemical properties of unsaturated FAME. Ab initio calculations for unsatu-
rated FAME will not only improve mechanisms for saturated FAME, but they will help
build comprehensive mechanisms for real biodiesel, which is a mixture of saturated and
unsaturated FAME.
Future experimental and modeling research should be directed towards unsaturated
FAME because current biodiesel fuels from soybean and canola oil, as well as next gen-
eration biodiesel from microalgal oil, have a high concentration of unsaturated FAME.
As shown in Figure 9.1 the unsaturated FAME typically found in biodiesel have double
bonds far away from the carbonyl group (e.g. methyl 9-octadecenoate, methyl 9,12-
octadecadienoate). However, as shown in this study, low molecular weight unsaturated
methyl esters with double bonds near the carbonyl group (i.e., methyl 2-propenoate,
methyl 3-butenoate, etc.) are important combustion intermediates. Therefore, this thesis
study recommends that future research be directed towards understanding the individual
eects of i. double bonds which are near the carbonyl group by using methyl 2-propenoate
and methyl 2-octenoate as surrogate compounds, and ii. double bonds which are far the
carbonyl group using methyl 9-nonenoate as as a surrogate compound
6
6
These surrogate compounds are selected because they represent the types of molecules encountered
in FAME combustion and they are readily available from chemical suppliers.
Literature Cited
[1] E. Fisher, W. Pitz, H. Curran, and C. Westbrook, Detailed chemical kinetic mecha-
nisms for combustion of oxygenated fuels, Proceedings of the Combustion Institute,
vol. 28, no. 2, pp. 1579 1586, 2000.
[2] S. Gail, M. Thomson, S. Sarathy, and S. Syed, A wide-ranging kinetic modeling
study of methyl butanoate combustion, Proceedings of the Combustion Institute,
vol. 31, pp. 305311, 2007.
[3] W. Schwartz, C. McEnally, and L. Pfeerle, Decomposition and hydrocarbon
growth processes for esters in non-premixed ames, Journal of Physical Chemistry
A, vol. 110, no. 21, pp. 66436648, June 2006.
[4] S. Dooley, H. J. Curran, and J. M. Simmie, Autoignition measurements and a
validated kinetic model for the biodiesel surrogate, methyl butanoate, Combustion
and Flame, vol. 153, no. 1-2, pp. 232, April 2008.
[5] W. K. Metcalfe, S. Dooley, H. J. Curran, J. M. Simmie, A. M. El-Nahas, and M. V.
Navarro, Experimental and modeling study of C5H10O2 ethyl and methyl esters,
Journal of Physical Chemistry A, vol. 111, no. 19, pp. 40014014, May 2007.
[6] S. M. Walton, M. S. Wooldridge, and C. K. Westbrook, An experimental investi-
gation of structural eects on the auto-ignition properties of two C-5 esters, Pro-
ceedings of the Combustion Institute, vol. 32, no. Part 1, pp. 255262, 2009.
[7] A. Farooq, D. F. Davidson, R. K. Hanson, L. K. Huynh, and A. Violi, An ex-
perimental and computational study of methyl ester decomposition pathways using
shock tubes, Proceedings of the Combustion Institute, vol. 32, no. Part 1, pp. 247
253, 2009.
189
Literature Cited 190
[8] L. K. Huynh, K. C. Lin, and A. Violi, Kinetic modeling of methyl butanoate in
shock tube, Journal of Physical Chemistry A, vol. 112, no. 51, pp. 13 47013 480,
December 2008.
[9] A. M. El-Nahas, M. V. Navarro, J. M. Simmie, J. W. Bozzelli, H. J. Curran, S. Doo-
ley, and W. Metcalfe, Enthalpies of formation, bond dissociation energies and re-
action paths for the decomposition of model biofuels: Ethyl propanoate and methyl
butanoate, Journal of Physical Chemistry A, vol. 111, no. 19, pp. 37273739, May
2007.
[10] A. Osmont, M. Yahyaoui, L. Catoire, I. Goekalp, and M. T. Swihart, Thermochem-
istry of c-o, (co)-o, and (co)-c bond breaking in fatty acid methyl esters, Combustion
and Flame, vol. 155, no. 1-2, pp. 334342, October 2008.
[11] A. Osmont, L. Catoire, I. Goekalp, and M. T. Swihart, Thermochemistry of c-c
and c-h bond breaking in fatty acid methyl esters, Energy & Fuels, vol. 21, no. 4,
pp. 20272032, July-August 2007.
[12] A. Osmont, L. Catoire, and I. Goekalp, Thermochemistry of methyl and ethyl esters
from vegetable oils, International Journal of Chemical Kinetics, vol. 39, no. 9, pp.
481491, September 2007.
[13] C. J. Hayes and D. R. Burgess, Jr., Exploring the oxidative decompositions of
methyl esters: Methyl butanoate and methyl pentanoate as model compounds for
biodiesel, Proceedings of the Combustion Institute, vol. 32, no. Part 1, pp. 263270,
2009.
[14] L. K. Huynh and A. Violi, Thermal decomposition of methyl butanoate: Ab initio
study of a biodiesel fuel surrogate, Journal of Organic Chemistry, vol. 73, no. 1,
pp. 94101, January 2008.
[15] T. Vaughn, Hammill, H. M., M., and A. Marchese, Ignition delay of bio-ester fuel
droplets, SAE Technical Paper Series, vol. 2006-01-3302, 2006.
[16] K. HadjAli, M. Crochet, G. Vanhove, M. Ribaucour, and R. Minetti, A study of
the low temperature autoignition of methyl esters, Proceedings of the Combustion
Institute, vol. 32, no. Part 1, pp. 239246, 2009.
Literature Cited 191
[17] C. Westbrook, W. Pitz, and H. Curran, Chemical kinetic modeling study of the
eects of oxygenated hydrocarbons on soot emissions from diesel engines, Journal
of Physical Chemistry A, vol. 110, no. 21, pp. 69126922, June 2006.
[18] P. A. Glaude, W. J. Pitz, M. J. Thomson, and R. Pitz, Chemical kinetic model-
ing of dimethyl carbonate in an opposed-ow diusion ame, Proceedings of the
Combustion Institute, vol. 30, no. 1, pp. 1111 1118, 2005.
[19] J. P. Szybist, A. L. Boehman, D. C. Haworth, and H. Koga, Premixed ignition
behavior of alternative diesel fuel-relevant compounds in a motored engine experi-
ment, Combustion and Flame, vol. 149, no. 1-2, pp. 112128, April 2007.
[20] P. Pepiot-Desjardins, H. Pitsch, R. Malhotra, S. Kirby, and A. Boehman, Structural
group analysis for soot reduction tendency of oxygenated fuels, Combustion and
Flame, vol. 154, pp. 191208, 2008.
[21] S. M. Sarathy, S. Gaeil, S. A. Syed, M. J. Thomson, and P. Dagaut, A comparison
of saturated and unsaturated C4 fatty acid methyl esters in an opposed ow diusion
ame and a jet stirred reactor, Proceedings of the Combustion Institute, vol. 31, no.
Part 1, pp. 10151022, 2007.
[22] S. Gail, S. M. Sarathy, M. J. Thomson, P. Dievart, and P. Dagaut, Experimental
and chemical kinetic modeling study of small methyl esters oxidation: Methyl (e)-
2-butenoate and methyl butanoate, Combustion and Flame, vol. 155, no. 4, pp.
635650, December 2008.
[23] Anonymous, Draft Technical Report EPA420-P-02-001, United States Environ-
mental Protection Agency, Tech. Rep., 2002.
[24] S. Garner, R. Sivaramakrishnan, and K. Brezinsky, The high-pressure pyrolysis of
saturated and unsaturated c
7
hydrocarbons, Proceedings of the Combustion Insti-
tute, vol. 32, pp. 461467, 2009.
[25] G. Dayma, S. Gail, and P. Dagaut, Experimental and kinetic modeling study of
the oxidation of methyl hexanoate, Energy & Fuels, vol. 22, no. 3, pp. 14691479,
May-June 2008.
[26] P. Dagaut, S. Gail, and M. Sahasrabudhe, Rapeseed oil methyl ester oxidation over
extended ranges of pressure, temperature, and equivalence ratio: Experimental and
Literature Cited 192
modeling kinetic study, Proceedings of the Combustion Institute, vol. 31, no. Part
2, pp. 29552961, 2007.
[27] O. Herbinet, W. J. Pitz, and C. K. Westbrook, Detailed chemical kinetic oxidation
mechanism for a biodiesel surrogate, Combustion and Flame, vol. 154, no. 3, pp.
507528, August 2008.
[28] T. Lu and C. K. Law, Toward accommodating realistic fuel chemistry in large-
scale computations, Progress in Energy and Combustion Science, vol. 35, no. 2, pp.
192215, April 2009.
[29] K. Seshadri, T. Lu, O. Herbinet, S. B. Humer, U. Niemann, W. J. Pitz, R. Seiser,
and C. K. Law, Experimental and kinetic modeling study of extinction and ignition
of methyl decanoate in laminar non-premixed ows, Proceedings of the Combustion
Institute, vol. 32, no. Part 1, pp. 10671074, 2009.
[30] C. McEnally, U. Koylu, L. Pfeerle, and D. Rosner, Soot volume fraction and
temperature measurements in laminar nonpremixed ames using thermocouples,
Combustion and Flame, vol. 109, no. 4, pp. 701720, June 1997.
[31] R. J. Kee, F. M. Rupley, J. A. Miller, M. E. Coltrin, J. F. Grcar, E. Meeks, H. K.
Moat, A. E. Lutz, G. Dixon-Lewis, M. D. Smooke, J. Warnatz, G. H. Evans,
R. S. Larson, R. E. Mitchell, L. R. Petzold, W. C. Reynolds, M. Caracotsios, W. E.
Stewart, P. Glarborg, C. Wang, C. L. McLellan, O. Adigun, W. G. Houf, C. P. Chou,
S. F. Miller, P. Ho, P. D. Young, D. J. Young, D. W. Hodgson, M. V. Petrova, and
K. V. Puduppakkam. (2006) CHEMKIN release 4.1. San Diego, California.
[32] H. Curran, Rate constant estimation for C1 to C4 alkyl and alkoxyl radical decom-
position, International Journal of Chemical Kinetics, vol. 38, no. 4, pp. 250275,
April 2006.
[33] D. Healy, H. J. Curran, J. M. Simmie, D. M. Kalitan, C. M. Zinner, A. B. Barrett,
E. L. Petersen, and G. Bourque, Methane/ethane/propane mixture oxidation at
high pressures and at high, intermediate and low temperatures, Combustion and
Flame, vol. 155, pp. 441448, 2008.
Literature Cited 193
[34] W. Tsang and R. Hampson, Chemical kinetic databse for combustion chemistry:
Part I. Methane and related compounds, Journal of Physical Chemical Reference
Data, vol. 15, 1986.
[35] T. Lu and C. Law, A directed relation graph method for mechanism reduction,
Proceedings of the Combustion Institute, vol. 30, pp. 13331341, 2005.
[36] , On the applicability of directed relation graph to the reduction of reaction
mechanisms, Combustion and Flame, vol. 146, pp. 472483, 2006., 2006.
[37] C. K. Westbrook, W. J. Pitz, O. Herbinet, H. J. Curran, and E. J. Silke, A com-
prehensive detailed chemical kinetic reaction mechanism for combustion of n-alkane
hydrocarbons from n-octane to n-hexadecane, Combustion and Flame, vol. 156,
no. 1, pp. 181199, January 2009.
[38] P. Dagaut, M. Reuillon, M. Cathonnet, and D. Voisin, High-pressure oxida-
tion of normal-decemberane and kerosene in dilute conditions from low to high-
temperature, Journal de Chimie Physique et de Physico-Chimie Biologique, vol. 92,
no. 1, pp. 4776, January 1995.
[39] E. R. Ritter and J. W. Bozzelli, THERM: Thermodynamic property estimation
for gas phase radicals and molecule, International Journal of Chemical Kinetics,
vol. 23, pp. 767778, 1991.
[40] S. Benson, Thermochemical Kinetics, 2nd edition. Wiley, New York, 1976.
[41] R. Sumathi and W. H. G. Jr., Oxygenate, oxyalkyl and alkoxycarbonyl thermo-
chemistry and rates for hydrogen abstraction from oxygenates, Physical Chemistry
Chemical Physics, vol. 5, pp. 34023417, 2003.
[42] L. Tee, S. Gotoh, and W. Stewart, Molecular parameters for normal uids - lennard-
jones 12-6 potential, Industrial & Engineering Chemistry Fundamentals, vol. 5,
no. 3, p. 356, 1966.
[43] H. Wang and M. Frenklach, Transport properties of polycyclic aromatic hydrocar-
bons for ame modeling, Combustion and Flame, vol. 96, no. 1-2, pp. 163170,
January 1994.
Literature Cited 194
[44] P. Linstrom and W. Mallard. (2005) NIST chemistry webbook, NIST standard
reference database number 69. National Institute of Standards and Technology,.
Gaithersburg, Maryland, 20899. [Online]. Available: http://webbook.nist.gov
[45] A. McClellan, Tables of Experimental Dipole Moments. San Francisco: Freeman,
1963.
[46] S. Bohm and O. Exner, Prediction of molecular dipole moments from bond mo-
ments: testing of the method by DFT calculations on isolated molecules, Physical
Chemistry Chemical Physics, vol. 6, no. 3, pp. 510514, 2004.
[47] D. R. Lide, Ed., CRC Handbook of Chemistry and Physics, 87th Edition. Boca Ra-
ton, FL: Taylor and Francis, 2007. [Online]. Available: http:/www.hbcpnetbase.com
[48] R. Bosque and J. Sales, Polarizabilities of solvents from the chemical composition,
Journal of Chemical Information and Computer Sciences, vol. 42, no. 5, pp. 1154
1163, September-October 2002.
[49] J. M. Simmie and H. J. Curran, Energy barriers for the addition of h, CH3, and
C2H5 to CH2=CHX [X = H, CH3, OH] and for h-atom addition to RCH=O [R = H,
CH3, C2H5, n-C3H7]: Implications for the gas-phase chemistry of enols, Journal
of Physical Chemistry A, vol. 113, no. 27, pp. 78347845, July 2009.
[50] G. Black, H. Curran, S. Pichon, J. Simmie, and V. Zhukov, Bio-butanol: Combus-
tion properties and detailed chemical kinetic model, Combustion and Flame, vol.
in press, p. doi:10.1016/j.combustame.2009.07.007, 2009.
[51] G. Dayma, C. Togbe, and P. Dagaut, Detailed Kinetic Mechanism for the Oxidation
of Vegetable Oil Methyl Esters: New Evidence from Methyl Heptanoate, Energ.
Fuel, vol. 23, pp. 42544268, SEP 2009.
[52] S. B. Dworkin, A. M. Schaer, B. C. Connelly, M. B. Long, M. D. Smooke,
M. A. Puccio, B. McAndrew, and J. H. Miller, Measurements and calculations of
formaldehyde concentrations in a methane/N-2/air, non-premixed ame: Implica-
tions for heat release rate, Proc. Combust. Inst., vol. 32, no. Part 1, pp. 13111318,
2009.
Literature Cited 195
[53] M. S. Kurman, R. H. Natelson, N. P. Cernansky, and D. L. Miller, Preignition
oxidation chemistry of the major jp-8 surrogate component: n-dodecane, American
Institute of Aeronautics and Astronautics, pp. 119, 2009.
Part IV
Closing
196
Chapter 10
Scientic Contribution
This dissertation researched the combustion kinetics of biofuels used in transportation
applications. The chemical compounds (i.e., oxygenates) in biofuels are structurally
dierent than those in conventional hydrocarbon fuels, so this study determined the role
of molecular structure on biofuel combustion and emissions. Chemical kinetic mechanism
were used to describe the molecular level transformation of reactants into products via
a series of elementary steps. Fundamental combustion experiments were performed, and
the acquired data was used to validate chemical kinetic mechanisms for biobutanol (i.e.,
n-butanol) and biodiesel (i.e., FAME). In addition, since biobutanol is a new biofuel
that has not been critically assessed for sustainability, this dissertation also performed
an LCA of biobutanol.
Biofuel combustion was primarily studied using experimentally measured species and
temperature proles in an opposed-ow diusion ame. Additionally, JSR species proles
and laminar ame speed data
1
were used to study biofuel combustion. Comprehensive
chemical kinetic mechanisms for the combustion of n-butanol and FAME were developed,
and then validated using the experimental data. The eects of molecular structure on
combustion were elucidated using the chemical kinetic mechanism to identify the domi-
nant routes of reactant consumption and product formation.
The experimental results indicate that the combustion of n-butanol and FAME pro-
duced appreciable quantities of carbon dioxide, carbon monoxide, water, methane, and
ethylene. Therefore, the emissions proles are similar to those of conventional hydro-
carbon fuels. The combustion of biofuels produces higher levels of oxygenated product
1
These experiments were performed by collaborating researchers.
197
Chapter 10. Scientific Contribution 198
species, such as aldehydes and ketones. These species sequester carbon in carbon-oxygen
bonds, and therefore prevent carbon from forming soot precursors (e.g., acetylene). How-
ever, these oxygenated species may have adverse environmental and health impacts that
warrants further investigation.
The chemical kinetic mechanisms developed in this study correlates well with the
experimental data presented herein. The modeling results indicated that the oxygenated
molecules in biofuels follow similar combustion pathways to the hydrocarbons in petroleum
fuels. The mechanisms suggest that oxygenated product species are formed directly from
the oxygenated moieties in the parent biofuel molecules (i.e., the original carbon-oxygen
bonds are preserved through the combustion process).
The detailed chemical kinetic mechanisms developed in this thesis have been vali-
dated against a range of experimental data; however, more data is required to test the
mechanisms across a wider range of temperatures, pressures, mixing conditions, and
mixture fractions. Ultimately, these mechanisms will be an important tool for studying
combustion in practical applications, such as internal combustion engines. Additional
work is needed to reduce the complexity of the mechanisms in order to model multi-
dimensional systems; however, the detailed mechanisms herein provide the fundamental
scientic basis for such reduced mechanisms.
Appendix A
Schematic of the opposed-ow diusion ame setup. Copyright c 2009 by Tim Chan.
N
2
O
2
C
o
m
p
r
e
s
s
e
d

n
i
t
r
o
g
e
n

s
u
p
p
l
y
C
o
m
p
r
e
s
s
e
d

o
x
y
g
e
n

s
u
p
p
l
y
F
l
o
w

m
e
t
e
r
C
o
m
p
r
e
s
s
e
d

a
i
r

(
f
r
o
m

b
u
i
l
d
i
n
g
)
F
u
e
l
(
M
a
s
t
e
r
F
l
e
x
)
P
u
m
p

d
r
i
v
e
A
t
o
m
i
z
e
r
B
u
r
n
e
r
(
T
e
l
e
d
y
n
e

H
a
s
t
i
n
g
s
)
M
a
s
s

f
l
o
w

m
e
t
e
r

&

c
o
n
t
r
o
l
l
e
r
H
e
a
t
i
n
g

t
a
p
e
M
a
s
s

f
l
o
w

m
e
t
e
r

&

c
o
n
t
r
o
l
l
e
r
E
v
a
p
o
r
i
z
e
r
O
p
p
o
s
e
d
-
F
l
o
w

D
i
f
f
u
s
i
o
n

F
l
a
m
e

S
e
t
u
p
C
o
p
y
r
i
g
h
t

2
0
0
9

b
y

T
i
m

C
h
a
n
199
Appendix B
This Appendix contains full-sized gures and complete datasets for the n-butanol com-
bustion study
200
Chapter 10. Scientific Contribution 201
The following gures are comparisons of the experimental and predicted concentration
proles obtained from the oxidation of n-butanol in a JSR at o=1, P=101.3 kPa, t=0.07s
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

CO
CO2
CH4
Chapter 10. Scientific Contribution 202
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C4H9OH
C3H7CHO
C3H6
1-C4H8
Chapter 10. Scientific Contribution 203
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6
Chapter 10. Scientific Contribution 204
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6
Chapter 10. Scientific Contribution 205
The following gures are comparisons of the experimental and predicted concentration
proles obtained from the oxidation of n-butanol in a JSR at o=2, P=101.3 kPa, t=0.07
s
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

CO
CO2
CH4
Chapter 10. Scientific Contribution 206
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C4H9OH
C3H7CHO
C3H6
1-C4H8
Chapter 10. Scientific Contribution 207
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6
Chapter 10. Scientific Contribution 208
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

H2O
H2
CH2O
O2
Chapter 10. Scientific Contribution 209
The following gures are comparisons of the experimental and predicted concentration
proles obtained from the oxidation of n-butanol in a JSR at o=0.5, P=101.3 kPa, t=0.07
s
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

CO
CO2
CH4
Chapter 10. Scientific Contribution 210
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C4H9OH
C3H7CHO
C3H6
1-C4H8
Chapter 10. Scientific Contribution 211
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6
Chapter 10. Scientific Contribution 212
1E-05
1E-04
1E-03
1E-02
1E-01
1E+00
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

H2O
H2
CH2O
O2
Chapter 10. Scientific Contribution 213
The following gures are comparisons of the experimental and predicted concentration
proles obtained from the oxidation of n-butanol in a JSR at o=0.25, P=101.3 kPa,
t=0.07 s
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

(
P
P
M
)
CO
CO2
CH4
Chapter 10. Scientific Contribution 214
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C4H9OH
C3H7CHO
C3H6
1-C4H8
Chapter 10. Scientific Contribution 215
1E-05
1E-04
1E-03
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

C2H2
C2H4
CH3CHO
C2H6
Chapter 10. Scientific Contribution 216
1E-05
1E-04
1E-03
1E-02
1E-01
1E+00
750 850 950 1050 1150 1250
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n

H2O
H2
CH2O
O2
Chapter 10. Scientific Contribution 217
The following gures are comparisons of the experimental and predicted concentration
proles obtained from the oxidation of n-butanol in a JSR at o=1.0, P=1013 kPa, t=0.7
s
1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
CO old
CO new
CO2 old
CO2 ew
CH4 old
CH4 new
Chapter 10. Scientific Contribution 218
1E-06
1E-05
1E-04
1E-03
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C4H8 old
C4H8 new
C4H9OH old
C4H9OH new
C3H7CHO old
C3H7CHO new
C3H6 old
C3H6 new
Chapter 10. Scientific Contribution 219
1E-06
1E-05
1E-04
1E-03
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C2H2 old C2H2 new
C2H4 old C2H4 new
CH3CHO old CH3CHO new
C2H6 old C2H6 new
Chapter 10. Scientific Contribution 220
1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
H2O old
H2O new
H2 old
H2 new
CH2O old
CH2O new
Chapter 10. Scientific Contribution 221
The following gures are comparisons of the experimental and predicted concentration
proles obtained from the oxidation of n-butanol in a JSR at o=2.0, P=1013 kPa, t=0.7
s
1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
CO old
CO new
CO2
CH4
CO2 new
CH4 new
Chapter 10. Scientific Contribution 222
1E-06
1E-05
1E-04
1E-03
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C4H8 old
C4H8 new
C4H9OH old
C4H9OH new
C3H7CHO old
C3H7CHO new
C3H6 old
C3H6 new

Chapter 10. Scientific Contribution 223
1E-06
1E-05
1E-04
1E-03
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C2H2 old C2H2 new
C2H4 old C2H4 new
CH3CHO old CH3CHO new
C2H6 old C2H6 new
Chapter 10. Scientific Contribution 224
1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
H2O old
H2O new
H2 old
H2 new
CH2O old
CH2O new
Chapter 10. Scientific Contribution 225
The following gures are comparisons of the experimental and predicted concentration
proles obtained from the oxidation of n-butanol in a JSR at o=0.5, P=1013 kPa, t=0.7
s
1E-06
1E-05
1E-04
1E-03
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C4H8 old
C4H8 new
C4H9OH old
C4H9OH new
C3H7CHO old
C3H7CHO new
C3H6 old
C3H6 new
Chapter 10. Scientific Contribution 226
1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
CO old
CO new
CO2 old
CO2 new
CH4 old
CH4 new
Chapter 10. Scientific Contribution 227
1E-06
1E-05
1E-04
1E-03
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
C2H2 old C2H2 new
C2H4 old C2H4 new
CH3CHO old CH3CHO new
C2H6 old C2H6 new
Chapter 10. Scientific Contribution 228
1E-06
1E-05
1E-04
1E-03
1E-02
750 850 950 1050 1150
Temperature (K)
M
o
l
e

F
r
a
c
t
i
o
n
H2O old
H2O new
H2 old
H2 new
CH2O old
CH2O new
Chapter 10. Scientific Contribution 229
The following gures are comparisons of the experimental and predicted concentration
proles obtained from the oxidation of n-butanol in an atmospheric opposed-ow ame
(5.89% C
4
H
9
OH, 42% O
2
).
0
200
400
600
800
1000
1200
1400
1600
1800
2000
0 5 10 15 20
DISTANCE FROM FUEL PORT (mm)
T
E
M
P
E
R
A
T
U
R
E

(
K
)
measured
corrected
predicted
Chapter 10. Scientific Contribution 230
0%
2%
4%
6%
8%
10%
12%
0 2 4 6 8 10 12 14 16 18 20
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
%
)
CO2
CO
C4H9OH
Chapter 10. Scientific Contribution 231
0
2000
4000
6000
8000
10000
12000
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
p
p
m
)
C2H4
C2H2
CH4
Chapter 10. Scientific Contribution 232
0
500
1000
1500
2000
2500
3000
3500
4000
4500
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
p
p
m
)
C2H6
C4H8
C3H6
Chapter 10. Scientific Contribution 233
0
50
100
150
200
250
300
350
400
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
p
p
m
)
C3H7CHO
C3H8
C3H4
1,3-C4H6
Chapter 10. Scientific Contribution 234
0
200
400
600
800
1000
1200
1400
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
p
p
m
) CH2O
CH3CHO
Chapter 10. Scientific Contribution 235
The following gures are the experimental concentration proles obtained from the
oxidation of n-butane in an atmospheric opposed-ow ame (5.89% C
4
H
9
OH, 42% O
2
).
0%
2%
4%
6%
8%
10%
12%
0 2 4 6 8 10 12 14 16
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
%
) CO2
CO
C4H10
Chapter 10. Scientific Contribution 236
0
1000
2000
3000
4000
5000
6000
7000
8000
9000
10000
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C2H4
C2H2
CH4
Chapter 10. Scientific Contribution 237
0
200
400
600
800
1000
1200
1400
1600
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C2H6
C4H8
C3H6
Chapter 10. Scientific Contribution 238
0
50
100
150
200
250
300
350
400
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
CH2O
CH3CHO
Chapter 10. Scientific Contribution 239
0
10
20
30
40
50
60
70
80
0 2 4 6 8 10 12
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C3H7CHO
C3H8
Appendix C
This Appendix contains full-sized gures and complete datasets for the methyl decanoate
combustion study.
240
Chapter 10. Scientific Contribution 241
The following gures are comparisons of experimental and computed proles obtained
from the oxidation of MD in an atmospheric opposed-ow ame (1.8% MD, 42% O
2
).
0
200
400
600
800
1000
1200
1400
1600
1800
2000
0 2 4 6 8 10 12 14 16 18 20
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
measured
corrected
predicted
Chapter 10. Scientific Contribution 242
0%
2%
4%
6%
8%
10%
0 2 4 6 8 10 12 14 16 18 20
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
%
)
CO2
CO
MD
Chapter 10. Scientific Contribution 243
0
2000
4000
6000
8000
10000
12000
14000
0 2 4 6 8 10
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C2H4
C2H2
CH4
Chapter 10. Scientific Contribution 244
0
200
400
600
800
1000
1200
1400
1600
1800
2000
0 2 4 6 8 10
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C2H6
1-C4H8
C3H6
C8H16
Chapter 10. Scientific Contribution 245
0
100
200
300
400
500
600
700
800
900
1000
0 2 4 6 8 10
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C2H4O
CH2O
CH2CO
Chapter 10. Scientific Contribution 246
0
50
100
150
200
250
300
350
0 2 4 6 8 10
DISTANCE FROM FUEL PORT (mm)
M
O
L
A
R

C
O
N
C
E
N
T
R
A
T
I
O
N

(
P
P
M
)
C3H4
C5H10
C6H12
C7H14
1,3-C4H6

Vous aimerez peut-être aussi