Sunteți pe pagina 1din 63

Progress in Energy and Combustion Science 44 (2014) 40e102

Contents lists available at ScienceDirect

Progress in Energy and Combustion Science


journal homepage: www.elsevier.com/locate/pecs

Review

Alcohol combustion chemistry


S. Mani Sarathy a, *, Patrick Oßwald b, Nils Hansen c, Katharina Kohse-Höinghaus d
a
Clean Combustion Research Center, King Abdullah University of Science and Technology, Thuwal 23955-6900, Saudi Arabia
b
Institute of Combustion Technology, German Aerospace Center (DLR), Pfaffenwaldring 38-40, D-70569 Stuttgart, Germany
c
Combustion Research Facility, Sandia National Laboratories, Livermore, CA 94551, USA
d
Chemistry Department, Bielefeld University, Universitätsstraße 25, Bielefeld D-33615 Germany

a r t i c l e i n f o a b s t r a c t

Article history: Alternative transportation fuels, preferably from renewable sources, include alcohols with up to five or
Received 22 December 2013 even more carbon atoms. They are considered promising because they can be derived from biological
Accepted 14 April 2014 matter via established and new processes. In addition, many of their physical-chemical properties are
Available online 14 June 2014
compatible with the requirements of modern engines, which make them attractive either as re-
placements for fossil fuels or as fuel additives. Indeed, alcohol fuels have been used since the early years
Keywords:
of automobile production, particularly in Brazil, where ethanol has a long history of use as an automobile
Biofuel
fuel. Recently, increasing attention has been paid to the use of non-petroleum-based fuels made from
Combustion chemistry
Alcohols
biological sources, including alcohols (predominantly ethanol), as important liquid biofuels. Today, the
Kinetic modeling ethanol fuel that is offered in the market is mainly made from sugar cane or corn. Its production as a
Pollutant emissions first-generation biofuel, especially in North America, has been associated with publicly discussed
Internal combustion engines drawbacks, such as reduction in the food supply, need for fertilization, extensive water usage, and other
Flame speed ecological concerns. More environmentally friendly processes are being considered to produce alcohols
Ignition delay from inedible plants or plant parts on wasteland. While biofuel production and its use (especially ethanol
and biodiesel) in internal combustion engines have been the focus of several recent reviews, a dedicated
overview and summary of research on alcohol combustion chemistry is still lacking. Besides ethanol,
many linear and branched members of the alcohol family, from methanol to hexanols, have been studied,
with a particular emphasis on butanols. These fuels and their combustion properties, including their
ignition, flame propagation, and extinction characteristics, their pyrolysis and oxidation reactions, and
their potential to produce pollutant emissions have been intensively investigated in dedicated experi-
ments on the laboratory and the engine scale, also emphasizing advanced engine concepts. Research
results addressing combustion reaction mechanisms have been reported based on results from pyrolysis
and oxidation reactors, shock tubes, rapid compression machines, and research engines. This work is
complemented by the development of detailed combustion models with the support of chemical kinetics
and quantum chemistry. This paper seeks to provide an introduction to and overview of recent results on
alcohol combustion by highlighting pertinent aspects of this rich and rapidly increasing body of infor-
mation. As such, this paper provides an initial source of references and guidance regarding the present
status of combustion experiments on alcohols and models of alcohol combustion.
Ó 2014 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2. Alcohol fuels e origins, sustainability, properties, and present use . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.1. Origins and sustainability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.2. Alcohol fuel properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.3. Present use of alcohol fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

* Corresponding author. Tel.: þ966 2 808 4626 (work), þ966 (0) 544 700 142
(mobile).
E-mail address: Mani.Sarathy@kaust.edu.sa (S.M. Sarathy).

http://dx.doi.org/10.1016/j.pecs.2014.04.003
0360-1285/Ó 2014 Elsevier Ltd. All rights reserved.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 41

List of abbreviations MON motor octane number


MP-VTST multi-path variational transition state theory
ARAS atomic resonance absorption spectrometry MS mass spectrometry
BDE bond dissociation energy MS-VTST multi-structure variational transition state theory
CCD charge-coupled device NDIR non-dispersive infrared
CI compression ignition NMR nuclear magnetic resonance
CMOS complementary metal oxide semiconductor NOx nitrogen oxides
CN cetane number NTC negative temperature coefficient
CRDS cavity ring down spectroscopy NUIG National University of Ireland Galway
CRV constrained reaction volume PAH polycyclic aromatic hydrocarbon
DI direct injection PDPA phase Doppler particle anemometry
EAD electronic absorption detection PEPICO photoelectron photoion coincidence spectrometer
EI electron ionization PES potential energy surface
EPR electron paramagnetic resonance PI photon ionization
EVC exhaust valve closing PIV particle image velocimetry
FID flame ionization detector PM particulate matter
FTIR Fourier transform infrared spectroscopy RA resonance absorption
GC gas chromatography RCM rapid compression machine
HC hydrocarbon REMPI resonance enhanced multiphoton ionization
HCCI homogenous charge compression ignition RF resonance fluorescence
HPLC high-pressure liquid chromatography RON research octane number
IR infrared RRKM RiceeRamspergereKasseleMarcus
JSR jet-stirred reactor SAR structureeactivity relationship
LDV laser Doppler velocimetry SMPS scanning mobility particle sizer
LHV lower heating value SI spark-ignition
LIF laser-induced fluorescence ST shock tube
LS laser schlieren TCD thermal conductivity detector
LTC low-temperature combustion UV ultraviolet
mb/d million barrels of oil per day VUV vacuum ultraviolet
MBMS molecular beam mass spectrometry YSI yield sooting idex
mboe/d million barrels of oil equivalent per day

3. Fundamental combustion experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46


3.1. Combustion properties in selected idealized experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2. Developments in experimental combustion research on alcohols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.3. Methanol, ethanol, and isomers of propanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.4. Butanol isomers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.5. Higher alcohols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4. Chemical kinetics of alcohol combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.1. The molecular structure of alcohols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2. Formulation of comprehensive chemical kinetic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.3. Reaction mechanism, reaction classes, and rate rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.4. Model testing and analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.5. The combustion chemistry of methanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.6. The combustion chemistry of ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.7. The combustion chemistry of propanol isomers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.8. The combustion chemistry of butanol isomers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.9. The combustion chemistry of higher alcohols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5. Insights into engine-relevant phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6. Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

1. Introduction increasing mobility of that population, the increasing production


and consumption of goods, and the transportation of those goods
Energy is among the most important factors influencing eco- are considerable concerns in future energy strategies. The need for
nomic and industrial development. Today, combustion of fossil clean, efficient, and affordable energy sources is often seen in
fuels accounts for more than two-thirds of the world’s primary contrast to the environmental and health issues connected with
energy utilization [1]. The growing global population, the fossil fuel use. Projections for the next two decades assume the
42 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

continued leading role of fossil fuel-based energy with a global Table 1


share of about 30% of primary energy to be needed and utilized in World fuel ethanol production in 2012 by region and country.

Asia, mainly China and India [1]. While many alternatives to coal, Region Millions of Gallons
petroleum, and natural gas are currently used or are being dis- North & Central America 13,768
cussed for power generation (e.g., hydroelectricity and nuclear, South America 5,800
solar, wind, geothermal, tidal or wave energy) transportation of Europe 1,139
people and goods depends on the availability of high-density en- Asia 952
Africa 38.31
ergy sources, primarily delivered to the market today as liquid fuels
derived from petroleum. Country Millions of Gallons
Because of associated greenhouse gas emissions, fossil fuel USA 13,300
combustion has been identified as a key factor in climate change. As Brazil 5,577
one example, Fig. 1 shows the sources of greenhouse gas emissions China 555
Canada 449
in the European Union (EU-27) in 2010. Around 80% came from
Australia 87.2
energy-related sources, with about 63% from direct combustion of
Data from Ref. [10].
fuels, including about 20% from fuel combustion for transportation
[2].
Emissions from petroleum combustion have also been impli-
International Energy Agency [8] suggest that biofuel use will triple
cated in the deterioration of air quality and, subsequently, of human
from 1.3 million barrels of oil equivalent per day (mboe/d) in 2010
health due to the formation of pollutants such as nitrogen oxides
to 4.5 mboe/d in 2035, driven primarily by blending mandates, with
(NOx), carbon monoxide (CO), particulate matter (PM) and soot, and
the fuel ethanol supply rising from 1 mboe/d in 2010 to 3.4 mboe/d
gaseous products from incomplete combustion. Because the
in 2035. To put this in perspective, global oil demand in the year
transformation to the use of alternative energy sources for trans-
2035 is projected to reach 99.7 million barrels of oil per day (mb/d),
portation is especially difficult, energy-efficient, clean combustion
up from 87.4 mb/d in 2011 [9].
processes remain an important research topic. With advances in
This article presents a comprehensive review of the combustion
the development of engine combustion systems, such as homoge-
chemistry of alcohol fuels. The article begins by providing infor-
nous charge compression ignition (HCCI) or related strategies that
mation on the history of alcohol production and use. Sustainability
can adapt to a variety of fuels, the use of biofuels as alternative neat
issues are addressed and important alcohol fuel properties are
fuels or fuel additives is possible [3e7]. Energy independence,
presented. As a prerequisite and background for the necessity of
possible decreases in CO2, soot, and unburned hydrocarbon (HC)
detailed combustion chemistry research, the combustion of alco-
emissions, and other factors are driving this possibility for trans-
hols and alcohol fuel blends in engines is addressed, with a focus on
formation, with biomass-related fuel sources preferably grown on
emission characteristics. The experimental and theoretical tools for
wasteland along with the use of agricultural residue, food waste,
addressing combustion chemistry of fuels are highlighted to pro-
woody plant parts, or algae biomass presenting important topics of
vide an understanding of desired information and observable
discussion and research [3].
quantities. With the stage thus set, the article aims to give an
Ethanol is the most widely used biofuel today. More than 7.3
extensive overview of fundamental combustion chemistry by
billion gasoline-equivalent gallons of ethanol were used as gasoline
considering studies of alcohol fuels in shock tubes, rapid
additives in 2009 in the USA. The increase in the use of biomass to
compression machines, jet-stirred and flow reactors, and laminar
more than 4% of total primary energy utilization in the USA was
flames. The value of information from such experiments is critically
therefore mainly due to ethanol production [3]. The largest pro-
discussed. Subsequently, alcohol combustion chemistry is dis-
duction of fuel ethanol was in the USA, followed by Brazil (see
cussed by presenting a large body of experimental and modeling
Table 1). In 2010, these two countries accounted for about 86% of
investigations covering the combustion of C1eC5 alcohols, with a
global fuel ethanol production [3]. Recent projections from the

Fig. 1. Sources of greenhouse gases in the European Union, 2010.


Reprinted from [2].
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 43

special emphasis on ethanol and the butanol isomers. Ignition, leading to the eventual rise of petroleum fuels. Methanol (i.e., wood
flame propagation, qualitative species assignment, and full quan- alcohol), derived from destructive distillation of wood, and ethanol
titative speciation, depending on the specific combustion regime, were commonly used as lamp fuels prior to the 19th century. The
are considered. The combustion kinetics and development of a first internal combustion engines developed in the 19th century,
comprehensive mechanism relevant to alcohol combustion is pre- including those by Samuel Morey, Nikolaus Otto, and George
sented. The article closes by presenting a summary of the dis- Brayton, utilized ethanol as the fuel. Numerous engineers in the
tinguishing features of alcohol combustion chemistry and an early 20th century observed that ethanol and ethanol/gasoline
outlook toward future research in this area. blends had anti-knock properties. The industrialist Henry Ford
(founder of Ford Motor Company) and research engineer Charles
2. Alcohol fuels e origins, sustainability, properties, and Kettering (former Vice President of General Motors and former
present use President of the Society of Automotive Engineers) were strong
proponents of ethanol as a fuel for internal combustion engines in
Ancient civilizations converted grains to ethanol using the USA.
fermentation to produce low-alcohol-content wines and beers. The internal combustion engine designer Sir Harry R. Ricardo
Archaeologists have discovered primitive forms of distillation extensively studied the use of alcohol fuels in combustion en-
equipment dating back to antiquity [11], but these were used pri- gines. In the Introduction to his 1931 book, “The High-Speed
marily to purify plant essences for medicinal purposes [12] and to Internal Combustion Engine”, Ricardo [25] advocated the use of
make perfumes. Allchin [13] suggested that excavations in the In- alcohol fuels. He explained, “It is perfectly well known that
dian sub-continent together with ancient texts from the region alcohol is an excellent fuel, and there is little doubt, but that
provide evidence that distillation of alcoholic beverages dates back sufficient supplies could be produced within the tropical regions
to c. 500 BCE. However, the first written account [14] of boiling wine of the British Empire, yet little or nothing is being done to
and observing flammable vapors is found in the writings of Jabir ibn encourage its development.” Ricardo then presented numerous
Hayyan (c. 721e815 CE). Al-Kindi (c. 801e873 CE) later described the experimental results testing hydrocarbon fuels (e.g., paraffins,
distillation of wine in his Kitab al-Taraffuq fi al-‘itr (The Book of the aromatics, olefins, and naphthenes) and alcohol fuels in both
Chemistry of Perfume and Distillations) [14]. This early knowledge spark-ignition (SI) and compression ignition (CI) engines. His
of wine distillation by Arab chemists was transmitted to the West, pioneering early work on pre-ignition, which eventually led to
and Latin texts from the medical schools of Salerno (c. 1100 CE) the development of fuel octane ratings, showed that ethanol
mention the production of highly pure alcohol [15]. Up to the allowed for operation at a higher compression ratio due to its
present, distillation remains the industry standard for separation decreased pre-ignition propensity. Ricardo also noted that
and purification of alcohols. alcohol fuels had higher latent heats of vaporization and lower
Today, the primary source of ethanol continues to be microbial flame temperatures than petroleum-based ones, and that these
fermentation of sugars and starches. It is envisaged that in the properties could be exploited to increase power output and
future, lignocellulose will also be a source [4,7]. Historically, decrease thermal losses.
methanol was produced from the destructive distillation of pyro-
lyzed wood, but more recently it has been made from natural gas or 2.1. Origins and sustainability
coal via gasification to synthesis gas followed by methanol syn-
thesis. Industry-scale renewable routes to methanol exist in the This review of research on alcohol fuel combustion chemistry
form of replacing coal with biomass in gasification processes [4]. would be incomplete without at least a brief mention of the
The two isomers of propanol (n-propanol and iso-propanol) are intensely debated issues around biofuel sustainability. The highly
typically produced from petrochemicals, but novel biological routes cited articles by Farrell et al. [26] and Goldemberg [27] demon-
employing Escherichia coli are being explored at the laboratory strated that renewable ethanol fuel can be sustainably produced in
scale [16,17]. The production of butanol isomers from biomass many countries. More recently, researchers [28,29] showed that
feedstock dates back to 1861 when Louis Pasteur first documented indirect land-use change (i.e., deforestation of Amazonian forests)
the fermentation of sugars to n-butanol [18]. The early 1900s caused by biofuel plantations in Brazil needs to be critically
witnessed a large growth in n-butanol production when Chaim assessed to ensure sustainable production. A review on biofuels by
Weizmann, while attempting to produce n-butanol for synthetic Koh and Ghazoul provides a unique perspective on the environ-
rubber production, discovered that Clostridium acetobutylicum mental and societal impacts of biofuels [30]. The rapid policy-
could convert sugars into a mixture of acetoneen-butanoleethanol driven growth of biofuel use has led to serious environmental
(ABE) [18]. This production process reached commercial scales in and food security concerns. Current biofuel technologies compete
the 1930s and 1940s, after which it lost market share to the more with the food industry for feedstock, and the diversion of corn,
economically viable petroleum-derived n-butanol. Nigam and rice, and oilseeds to biofuel production is cited as the cause of
Singh [19] and Lee et al. [20] briefly reviewed the recent activities in rising food prices and global food shortages. In addition, large
bio-butanol research and indicated that biological pathways exist amounts of forested lands are being destroyed for biofuel feed-
for the production of n-butanol, iso-butanol, and 2-butanol [16,21]. stock production, leading to a loss in biodiversity and carbon-rich
The fourth isomer of butanol, tert-butanol, is a petrochemical sinks. The competition for fresh water resources presents an
product. The higher alcohols, including the isomers of pentanol and additional barrier to widespread biofuel use. Despite these chal-
hexanol, are currently derived from petrochemicals on relatively lenges, advances in biofuel feedstock and production technologies
small scales. The production of pentanol and hexanol isomers (e.g., may have the potential to ameliorate the negative impacts of
3-methyl-1-butanol, 2-methyl-1-butanol, 3-methyl-1-pentanol) biofuels.
from biological sources has become an area of recent research ac- It is clear from the literature that conventional bio-alcohol
tivity [16,22,23]. production from food crops, most notably corn, has limited po-
In their book “The Forbidden Fuel: A History of Power Alcohol”, tential [28,31]. Reasons for this include competition from the food
Bernton, Kovarik, Sklar et al. [24] provide a comprehensive history industry [32], limited agricultural land for crop growth [31], and
of the production and use of alcohol fuels (primarily ethanol) in the high energy input requirements for agricultural chemicals and
19th and 20th centuries, as well as the political and market factors harvesting [28]. For biofuels to live up to their potential as suitable
44 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

alternative fuels, lignocellulosic biomass must be considered. 2.2. Alcohol fuel properties
Agricultural and forest residues as well as sources of lignocellu-
losic biomass offer advantages over food crop feedstock. These Fuel property requirements for SI engine fuels are dictated by
advantages include the ability for lignocellulosic biomass to be ASTM D4814, Standard Specification for Automotive Spark-Ignition
cultivated on marginal agricultural land, lower agricultural Engine Fuels, and for CI engine fuels by ASTM D975, Standard
chemical requirements, lower energy requirements, and the po- Specification for Diesel Fuel Oils. The physical and chemical prop-
tential for utilizing the lignin portion of the biomass as an energy erties of several alcohol, gasoline, and diesel fuels are presented in
source for the production process. For these reasons, the projected Table 2. This limited data set includes some of the important
net energy gains from lignocellulosic alcohols are much higher properties relevant to utilization of alcohol fuels in combustion
than those from corn-based alcohol production technologies, and systems. The table includes all C1eC4 isomers, two C5 isomers, and
the resulting greenhouse gas emissions would then be much normal C6eC8 isomers of alcohols along with their molecular
lower. Despite these inherent advantages, the conversion of structures. The following conclusions can be drawn from Table 2.
lignocellulosic feedstock to fuels is not yet at commercial scale,
primarily because it is challenging to convert lignocellulosic  An increase in the molecular weight of alcohol fuels corresponds
biomass to alcohols without offsetting some of the possible net to a decrease in the oxygen content (wt%) and an increase in the
energy savings [28,31e33]. carbon and hydrogen content. This results in an increase in the

Table 2
Physical and chemical properties of alcohol, gasoline, and diesel fuels.

Fuel Structure MW O2 (wt%) LHVa (MJ/L) Stoichiometric Boiling RONd MONd DHvape (kJ/kg Solubility Specific gravityg,h
air/fuel ratiob Pointc ( C) from 25  C) in waterg at 20  C
at 25  C, wt%

Methanol 32.04 0.50 15.8 6.46 64.7 109 89 1168 Miscible 0.792

Ethanol 46.06 0.35 21.4 8.98 78 109 90 919.6 Miscible 0.794

n-Propanol 60.09 0.27 24.7 10.33 97 104 89 792.1 Miscible 0.804

iso-Propanol 60.09 0.27 24.1 10.33 83 117 95 756.6 Miscible 0.789

n-Butanol 74.11 0.22 26.9 11.17 118 98 85 707.9 7.4 0.81

2-Butanol 74.11 0.22 26.7 11.17 99 105 93 671.1 18.1 0.808

iso-Butanol 74.11 0.22 26.6 11.17 108 105 90 686.4 8.1 0.802

tert-Butanol 74.11 0.22 25.7 11.17 83 107 94 629.9 Miscible 0.789

n-Pentanol 88.14 0.18 28.5 11.74 138 80 74 647.1 2.2 0.816

3-Methyl-1-butanol 88.14 0.18 27.8 11.74 130 94 84 617.1 2.7 0.8

n-Hexanol 102.16 0.16 29.3 12.15 157 56 46 603.0 0.6 0.814

n-Heptanol 116.19 0.14 30.1 12.47 175 e e 575.0 0.2 0.822

n-Octanol 130.21 0.12 31.1 12.71 195 28 27 545.0 0.05 0.826

Gasoline Average C8H15 111.19 0.00 30e33 14.58 27e225 88e98 80e88 w351 Negligible 0.72e0.78
Diesel Average C14H30 198.4 0.00 35.66 14.95 125e400 <0 <0 w232f Negligible 0.81e0.89f
a
Lower heating values for methanol and diesel Agarwal [7]; values for tert-butanol, iso-pentanol, and larger alcohols are derived from DHC; all others from Christensen et al.
[37].
b
The mass basis air/fuel ratio is calculated using 28.97 g/mol as the average molecular weight of air and O2 composition of 21 mol%.
c
Gasoline boiling point range from Christensen et al. [37]; diesel from Agarwal [7]; all others are average values from a survey of the literature.
d
Volumetric blending research octane number (RON) and motor octane number (MON) values are from Christensen et al. [37] for n-pentanol and 3-methylbutanol; neat
RON and MON for methanol and ethanol are from Hunwartzen [39] and Bauer et al. [40], as presented in Anderson et al. [41]; neat RON and MON for tert-butanol and iso-
propanol is from Hamadi [42]; neat RON and MON for n-hexanol and n-octanol are from Graboski et al. [43]; neat RON and MON for n-propanol, n-butanol, 2-butanol, and iso-
butanol from Christensen et al. [37].
e
Heat of vaporization, DHvap, from Christensen et al. [37].
f
“Diesel Fuel Oils, 1987,” Petroleum Product Surveys, National Institute for Petroleum and Energy Research, October 1987.
g
CRC handbook of chemistry and physics, 87th edition. Handbook of Chemistry and Physics, 91st ed.; CRC Press: Boca Raton, FL, 2010/2011.
h
Specific gravity is defined as the ratio of the density of fuel to the density of water at 20  C.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 45

energy content, indicated by the fuel’s volumetric lower heating  An interesting feature of alcohols is their high sensitivity,
value (LHV). C1eC3 alcohols have significantly lower LHVs than defined as the research octane number minus the motor octane
gasoline and diesel. Use of C1eC3 alcohols therefore tends to number (S ¼ RON  MON). This unique fuel property can have
reduce fuel economy (i.e., miles per gallon). Alcohols with four benefits in modern LTC engines and highly boosted direct in-
or more carbons have LHVs closer to that of gasoline, and are jection spark ignition (DI SI) engines that prefer high sensitivity
thus beneficial for improving fuel economy [34]. fuels [38].
 The presence of the hydroxyl moiety in alcohols causes them to
be liquid at room temperature and thus suitable fuels for trans-
portation applications. The lower alcohols (C1eC2) have lower 2.3. Present use of alcohol fuels
boiling points, which depresses front-end distillation when
blended with gasoline and diesel fuels. At present, US regulations Given the documented history of the use of alcohol fuels since
fix ethanol concentrations in gasoline, so refiners vary butane the internal combustion engine was developed and their favorable
concentrations in the fuel to meet summer and winter front-end physicalechemical properties, it is no surprise that alcohols
distillation specifications. In conventional carbureted SI engines, continue to be attractive fuels for today’s modern SI and CI engines.
this often leads to the problem of vapor lock, but recent de- The purpose of this study is to highlight the importance of alcohol
velopments in fuel injection (e.g., port fuel and direct injection) combustion in engines and thus to motivate the full investigation of
are more tolerant of volatile fuels [34]. Nevertheless, altering the their combustion chemistry. The following articles give a compre-
distillation curve modifies fuel spray formation, vaporization, hensive overview of the use of alcohol fuel in engines.
and in-cylinder mixture stratification, which are important
design considerations for modern low-temperature combustion  Agarwal [7] reviewed a large body of literature to conclude that
(LTC) engines [35,36]. The C3eC8 alcohols tend to have boiling ethanol can be blended with gasoline fuel to improve engine
points in the mid-point of the gasoline/diesel fuel range; they performance and reduce emissions of CO and unburned HCs in
can be blended with petroleum fuels at greater concentrations exhaust. Up to 20% of ethanol can be blended with diesel fuel to
with lesser impact on the fuel distillation curve. reduce CO and NOx emissions. Hansen et al. [6] also discussed
 The latent heat of vaporization (DHvap) of alcohols is signifi- the potential uses of ethanol-diesel fuel blends, highlighting
cantly higher than those of petroleum gasoline and diesel fuels, important blending characteristics and safety considerations.
which negatively impacts the ability to start the engine under  Giakoumis et al. [44] studied diesel engines running on blends
cold conditions (i.e., cold-start). However, current advances in of ethanol or n-butanol with diesel fuel under transient condi-
fuel delivery systems are largely able to deal with this problem. tions. They concluded that alcohol/diesel blends decrease
The higher molecular weight alcohols have less difficulty with engine-out PM and CO emissions, while increasing unburned HC
cold-start operation due to their lower specific latent heat of and oxygenated (e.g., carbonyl) emissions. A clear trend on NOx
vaporization. A fuel’s latent heat of vaporization can also affect emissions was not observed.
in-cylinder thermal stratification, which is important for con-  n-Butanol was studied as a fuel or as a blending agent for use in
trolling excessive pressure rise rates in LTC engines [35,36]. SI engines [34,45e52], CI engines [46,53e62], and motored
 The hydroxyl moiety in alcohols introduces a dipole moment to engines [63]. iso-Butanol has also received attention from en-
the non-polar hydrocarbon backbone. C1eC3 alcohols have short gine researchers with studies performed on SI engines [64e67]
hydrocarbon chains, so they are more polar and more soluble in and in CI engines [68,69]. Niass et al. [70] demonstrated that 2-
water than non-polar hydrocarbons. This fact lowers the upper butanol and tert-butanol could be blended with gasoline fuel for
limit of blending in petroleum fuels without the use of a co- improved engine performance.
solvent. Their water affinity also introduces challenges in their  In a series of papers, Gautam et al. studied combustion [34] and
distribution infrastructure, which must prevent water contam- emission [47] characteristics of C1eC5 normal alcohols blended
ination. Higher alcohols are increasingly less polar due to their with gasoline. Their results indicated that there was substantial
longer non-polar hydrocarbon chains. They are thus easier to knock resistance when using alcohol fuels blends as well as an
blend with non-polar hydrocarbons and they have lower affinity ability to reach higher compression ratio operation.
for water.  Vancoillie et al. [71] recently demonstrated the use of methanol
 The ignition propensity (i.e., knock resistance, octane rating) of as a fuel for SI engines at diesel-like efficiencies (up to 50%) and
alcohols is perhaps their most attractive feature for internal emission levels comparable to or lower than those of gasoline.
combustion engine applications. A high octane rating correlates
with a lower propensity for ignition and allows the SI engine to In summary, these papers suggest that alcohols with various
operate at a higher compression ratio without knocking. molecular structures (i.e., carbon number and substitution) can be
Methanol, ethanol, and propanol isomers have high research used in today’s SI and CI engines with minor modifications to the
octane number (RON) and have long been used to improve the engine design. However, the details of the combustion chemistry of
octane number of SI engine fuels. Butanol isomers have octane alcohol fuels, including pathways to potential pollutants, are less
ratings similar to that of conventional gasoline and are thus understood than are those of the hydrocarbons that constitute to-
suitable for SI engines. C5- and larger alcohols have octane rat- day’s petroleum-derived transportation fuels [72]. Some aspects of
ings that are lower than that of gasoline, and hence decrease concern regarding their perceived “cleaner combustion” image are
knock resistance when blended with petroleum SI engine fuels the emissions of carbonyl species, such as formaldehyde (CH2O),
[37]. However, a decreased octane number correlates with an acetaldehyde (ethanal, C2H4O), acetone (C3H6O), acrolein (prope-
increased cetane number (i.e., greater propensity for ignition), nal, C3H4O). These compounds present an air quality concern
making the larger alcohols suitable for CI engine applications. because they are toxic and irritating, as well as being precursors to
Alcohol fuel branching such as methyl substitution tends to urban smog (e.g., free radicals, ozone. and peroxyacyl nitrates) [73e
decrease the ignition propensity (i.e., increase the octane num- 75]. Pang et al. [76] conducted carbonyl emissions measurements in
ber). Later sections of this paper provide fundamental expla- SI and CI engines using blends of ethanolegasoline and ethanole
nations of how chemical features such as hydroxyl moiety and biodieselediesel. Their results showed that ethanol blending
methyl substitution affect the ignition propensity. reduced tailpipe emissions of CO and PM, but increased total
46 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Fig. 2. Percentage of carbonyl compounds in engine-out emissions from E10 gasoline at six load levels under stable speed (2800 rpm). For (formaldehyde), Ace (acetaldehyde),
Arn (acrolein), Atn (acetone), Pro (propionaldehyde), Cro (crotonaldehyde), Mac (methylacrolein), But (butyraldehyde), Ben (benzaldehyde), Tol (m-tolualdehyde, p-tolualdehyde),
others include valdehyde, hexaldehyde, and 2,5-dimethylbenzaldehyde.
Reprinted from [76] with permission from Elsevier.

tailpipe carbonyl emissions. Engine-out emissions included high Reviewing several earlier studies, the authors suggested the level of
concentrations of acetaldehyde (e.g., see Fig. 2), which is a product aldehyde emission is a function of engine type, load, and com-
of ethanol combustion. bustion mode in addition to different pathways that may lead to
Armas et al. [77] showed that ethanol and butanol blends with their formation.
diesel could reduce PM during warm engine start, whereas com- Aldehyde formation and emission from combustion sources
bustion instabilities led to increases in unburned hydrocarbon and may cause underestimated health problems. A study by Jacobson
CO emissions under cold-start conditions. Similarly, addition of n- [79] concluded that widespread ethanol use may increase the risk
butanol was seen to decrease smoke density, CO, and NOx emis- of cancer and ozone-related illnesses due to higher aldehyde
sions in diesel-n-butanol blends [57], but it increased (not further emissions and increased unburned ethanol emissions, which break
specified) unburned hydrocarbon emissions. The use of increasing down to acetaldehyde in the atmosphere. Combustion chemistry
amounts of n-butanol in gasoline in HCCI engines was seen to studies of alcohols can help to determine the importance of
significantly influence formaldehyde and acetaldehyde emissions oxygenated emissions by elucidating the role of fuel-bound oxygen
in a study by He et al. [78]. Formaldehyde levels increased with during combustion.
increasing n-butanol addition, almost irrespective of engine speed. It is challenging to draw broad scientific conclusions from
The study concluded that formaldehyde, once formed, cannot be vehicle engine studies, especially in terms of emissions character-
further oxidized in the cylinder under the conditions tested ization, because of the highly variable nature of tested boundary
because n-butanol blending advanced ignition timing, reduced conditions across studies (i.e., engine-out versus tailpipe, after-
combustion duration, and decreased in-cylinder temperatures treatment technologies, varying loads, speeds, engine types,
during the exhaust stroke. Emission of acetaldehyde were nearly fueling methods, intake temperatures and pressures, and operating
independent of engine speed under the conditions investigated, conditions). Nevertheless, engine research demonstrates the utility
but at any given exhaust valve closure timing, acetaldehyde levels of alcohol fuels in practical applications and motivates the revision
increased with the amount of n-butanol in the blend (Fig. 3). Such of regulatory policies around fuel and emission standards. Work on
effects were suggested to be a result of the different reaction scientific approaches toward understanding the effects of alcohol
mechanisms of gasoline and n-butanol to form acetaldehyde [78]. fuels on combustion is the focus of the remainder of this paper,
wherein fundamental chemical kinetic combustion experiments
and simulations with well-defined boundary conditions are
described. The fundamental combustion properties to be reviewed
for alcohol fuels are the premixed laminar flame speed, homoge-
nous ignition delay time, and product species distributions from
gas-phase oxidation/combustion.

3. Fundamental combustion experiments

Combustion in internal combustion engines is a complex pro-


cess involving fuel atomization, vaporization, fuel-air mixing,
ignition, and combustion. For example, Dec [80] has presented a
phenomenological description, or “conceptual model”, for diesel
combustion in a direct injected (DI) CI engine. As shown in Fig. 4,
liquid fuel is injected as a high-velocity spray into the combustion
chamber, where it vaporizes upon impingement with high-
temperature, high-pressure in-cylinder gases. Low-temperature
reactions spontaneously ignite portions of premixed fuel and air
Fig. 3. The effect of exhaust valve closing (EVC) timing on acetaldehyde emissions with
at local equivalence rations of about 2e4 (i.e., fuel-rich conditions),
butanolegasoline blends. which results in rapid heat release. The remaining fuel spray is then
Reprinted from [79] with permission from Elsevier. consumed in a high-temperature diffusion flame. Combustion
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 47

models. Finally, we demonstrate some examples of utilizing


detailed chemical kinetic modeling simulations to understand
alcohol fuel effects in internal combustion engines.

3.1. Combustion properties in selected idealized experiments

Investigating the combustion chemistry of alcohols requires


conditions under which chemical kinetic processes govern the
conversion of the fuel and oxidizer into products, and where mix-
ing processes are considered negligible or at least limited only to
gas-phase diffusion (e.g., perfectly mixed or slow mixing). These
requirements have led researchers to the use of idealized reactors
and flames for studying the global gas-phase combustion chemistry
Fig. 4. A conceptual model for diesel combustion in a DI CI engine. of alcohols at various temperatures, pressures, and mixture com-
Reprinted from Dec [80] with permission from SAE International. positions. These include, but are not limited to, homogenous gas-
phase batch reactors (e.g., shock tubes and rapid compression
machines), jet-stirred reactors, plug flow reactors (e.g., laminar flow
gases are produced throughout the expansion process with insuf-
and turbulent flow reactors), laminar premixed flames, and non-
ficient oxygen available to completely oxidize all the intermediates.
premixed diffusion flames. The fundamental combustion proper-
This unsteady, heterogenous, three-dimensional process is chal-
ties investigated using these experimental systems include the
lenging to model, and it is difficult to decouple physical mixing
homogenous ignition delay time, premixed laminar flame speed,
processes from chemical kinetic processes. A similar conceptual
and product species distributions from gas-phase oxidation and/or
model has recently been presented for partially premixed LTC
pyrolysis. The respective property can be monitored as a function of
diesel engines under low-load conditions by Musculus, Miles, and
time t, temperature T, pressure P, or a specific spatial coordinate x.
Pickett [81].
It has been demonstrated that a fuel’s homogenous ignition
Computer simulations are capable of combining fluid dynamics,
delay can be correlated with its octane rating [86,87], which is a
spray dynamics, chemical kinetics, and heat and mass transfer to
critical parameter for engine design and operation. Kalghatgi [88]
reproduce ignition behavior, pollutant formation, energy release,
notes that octane ratings are primarily determined by chemical
and other features of engine operation. Such simulations are widely
kinetics rather than physical fuel effects (e.g., latent heat of vapor-
used in the automobile industry to increase fuel economy and
ization, density, viscosity, physical mixing effects, etc.). The pre-
reduce emissions. Westbrook et al. [82] provided a thorough review
mixed laminar flame speed is a fundamental indicator of fuel
of computational combustion methods and their applications to-
reactivity at high temperature, and is also a critical parameter for
ward simulation of combustion in SI, CI, and HCCI engines. Typi-
determining the turbulent flame speed in engines. The emissions
cally, these engine simulations are computationally expensive, so
formed upon fuel oxidation and/or pyrolysis in ideal reactors are
simplifications in fluid dynamics, spray dynamics, and elementary
indicative of the species formed in engines under similar temper-
chemical kinetics are made to run them on present personal
ature, pressure, and equivalence ratio conditions. Thus, under-
computing technologies.
standing combustion in idealized systems can offer insights into a
Reducing the chemical fidelity in simulations would limit our
fuel’s performance in practical engines.
ability to understand the comprehensive combustion chemistry of
It is important to note that several researchers have had suc-
alcohols. Recent studies by Westbrook et al. [83,84] have shown
cess studying alcohol combustion chemistry in non-idealized
that an understanding of engine-relevant phenomena can be ob-
experimental setups. For example, Haas et al. [89] used an igni-
tained using detailed chemical kinetic models in simpler reacting
tion quality tester to study the ignition chemistry of butanol iso-
flow simulations, given the appropriate boundary conditions. For
mers in reacting sprays. By matching the physical properties of the
example, the effects of oxygenated hydrocarbons on soot emissions
fuels, they were able to attribute differences in ignition delay
from diesel engines were studied [83] using detailed kinetic
times to chemical kinetic processes rather than physical processes
modeling simulations of idealized rich premixed ignition and
(e.g., spray formation and droplet vaporization). In addition, HCCI
combustion of diesel surrogate fuels. The boundary conditions (e.g.,
engines have been used to study the combustion chemistry of
temperature, pressure, and equivalence ratios) were obtained from
alcohol fuels with reasonable levels of success. Some notable
Dec’s conceptual model for DI CI diesel engine combustion [80]. The
studies include investigations on interesting features such as HCCI
relative sooting tendencies of various fuels were then determined
engine low- and intermediate-temperature heat release and
by summing the concentrations of soot precursors (e.g., acetylene,
combustion phasing for ethanol [59,90e95], n-butanol [59,96],
benzene, ethylene, toluene, etc.) following the rich ignition event.
and iso-pentanol [97]. These studies offer insights into alcohol
In separate studies, Westbrook has shown that similar simulation
fuel/engine interactions, and describe engine-related phenomena
approaches can be used to the predict gasoline fuel octane ratings
that could be better understood with fundamental combustion
[85], HCCI engine pressure traces [85], and biodiesel fuel cetane
experiments at relevant temperature, pressure, and mixture
ratings [84].
fraction conditions.
The success of the aforementioned approach toward improving
The homogenous ignition delay is the time it takes for a ho-
an understanding of fuel effects in engines motivates us to first
mogenous gas-phase fuel/air mixture at a specified temperature
describe the study of alcohol fuel in idealized combustion systems
and pressure to reach an explosive state resulting in a rapid release
(e.g., reactors and flames). In this way, the fuel’s combustion
of heat and a rise in temperature and pressure. Ignition delay time
chemistry can be precisely modeled, such that the effects of the
measurements for fuel/air mixtures at high pressures (e.g., greater
molecular structure of the alcohol on combustion and emissions
than 10 bar) and low to intermediate temperatures (e.g., 600e
can be understood. We then delve deeper into the chemical kinetics
1000 K) tend to correlate well with fuel octane and cetane ratings,
and combustion chemistry mechanisms of alcohols, including
and thus are the most relevant to engine operating conditions.
simulation of idealized reactors using detailed chemical kinetic
Fuels with longer ignition delay times under the aforementioned
48 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

conditions tend to have higher octane numbers and lower cetane of low-temperature ignition unique to alcohol fuels are described in
numbers. Ignition delay time measurements at lower pressures, Section 4.
higher temperatures, and/or on highly diluted mixtures are not The premixed laminar flame speed is simply defined as the
representative of real engine conditions, but still provide valuable velocity at which a flame front propagates into an unburned gas
data for the testing and/or validation of chemical kinetic models. mixture at a fixed temperature, pressure, and mixture composition.
Recent reviews on low-temperature ignition are available in this It is a fundamental combustion property that is employed in
journal by Battin-Leclerc [98] and by Zádor et al. [99]. Battin-Leclerc describing more complex combustion phenomena existing in en-
[98] presents a comprehensive description of various experimental gines, such as the turbulent flame speed, flame structure, flame
setups used for measuring ignition delay times. For alcohols, these stabilization and instabilities, and flame extinction [109]. The
properties have been largely measured using rapid compression laminar flame speed is an idealized combustion property that
machines (RCMs) and shock tubes (STs). In both experiments, a test cannot be measured exactly. Therefore, experiments measuring
gas mixture comprising fuel, O2, and a diluent (e.g., N2, Ar, etc.) is laminar flame speeds attempt to achieve near-ideal conditions and
rapidly raised to the desired temperature and pressure, and then account for non-ideal effects, such that they can be subtracted from
held under near-constant-volume conditions until an ignition the experimental data [109]. Laminar flame speed measurements of
event occurs, typically measured as a rapid pressure rise or alcohols have been performed using a variety of experimental ar-
chemiluminescence emission. rangements, including counterflow diffusion flames, spherically
Fig. 5 presents typical ignition delay time measurements ob- propagating flames, and burner-stabilized flames. The propagating
tained in shock tubes and RCMs. When interpreting and modeling flame front is typically observed using direct photography of the
shock tube data, it is important to consider how the measurements luminous zone, shadowgraphy, or schlieren imaging. A schematic
were performed [100,101] and the existence of pre-ignition pres- of the counterflow twin-flame configuration at the University of
sure rise [102,103]. Two approaches exist for handling pre-ignition Southern California [110], which has been used to measure the
pressure rise in zero-dimensional homogenous batch reactor sim- laminar flame speeds of many C1eC5 alcohols, is shown in Fig. 6.
ulations: the volume-varying approach and the pressure-varying The effects of the molecular structure of alcohol fuels on laminar
approach. In the former method, the experimentally measured flame speeds are discussed in Section 4.
pressure time-history is converted to a volume history [102] Emissions are undoubtedly a chief concern in the use of com-
(assuming isentropic compression) for use in the simulation, bustion technologies. Correlating the kind and quantity of
whereas in the latter approach the measured pressure profile up to combustion-generated species with the molecular structure of the
the ignition time is used in the simulation to constrain the pressure. fuel again requires experiments with idealized flames and reactors.
RCM data also need to be carefully interpreted by evaluating the The distribution of species depends on the type of experimental
compression time, compression ratio, heat losses before and after apparatus and its operating temperature, pressure, and mixture
the end of compression [104e106], and potential reactions during composition. Species generated from alcohol fuel combustion have
the compression stroke. Ignition is a transient process of fuel been studied in shock tubes, RCMs, premixed flames, non-
reacting with oxygen to form intermediate species that eventually premixed diffusion flames, flow reactors, and stirred reactors. The
result in a radical chain-branching process [107]; experiments in species within the reacting gas and/or burnt gas mixture are
stirred reactors and flow reactors that measure the transient and/or analyzed online by using optical diagnostics or offline by trans-
steady-state formation and destruction of low- and intermediate- ferring gases into a chemical analyzer. Online diagnostic techniques
temperature species can also aid in understanding the ignition do not disturb the idealized flow conditions in the experiment, and
chemistry of alcohols. Further details on the chemical mechanisms they preserve the chemical nature of the test gas being sampled.

Fig. 5. (Left) shock tube ignition delay time measurements performed at Stanford University on 2-butanol in 4% O2 diluted in Ar, f ¼ 1, T ¼ 1176 K, P ¼ 0.5 bar. Reprinted from
Stranic et al. [108] with permission from Elsevier. (Right) RCM ignition delay measurements performed at University of Connecticut on 2-butanol. P(t) is the pressure as a function of
time, and P0 (t) is the time derivative of the pressure, as a function of time.
Reprinted from Weber and Sung [105] with permission from Elsevier.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 49

Fig. 6. Schematic of the experimental counterflow configuration used to measure laminar flame speeds of alcohol fuels at the University of Southern California.
Reprinted from Ref. [110] with permission from Elsevier.

These methods can measure a limited number of species present at reactive zone and/or sampled gases. In combination with
relatively high concentrations using a variety of absorption and synchrotron-based vacuum ultraviolet (VUV) single photon ioni-
emissions spectroscopy techniques in the UV to IR wavelength zation, the MBMS technique (PI-MBMS) is even able to provide
ranges [111,112]. Offline techniques offer the ability to measure a isomer-resolved quantitative species information from reactive
larger number of species at lower concentrations than offered by environments such as the reaction zone of a flame [118,119]. Most
online techniques, but at the expense of perturbing the experiment recently, the technique was extended by the successful combi-
and test gas during the sampling process. The test gas is often nation of a molecular-beam flame-sampling setup to a photo-
extracted to a gas chromatograph with various detectors (e.g., flame electron photoion coincidence spectrometer (PEPICO) [120]. This
ionization detector, thermal conductivity detector, mass spec- technique adds an additional analytical dimension due to the
trometer) or infrared spectrometers (e.g., Fourier transform observation of molecule-specific transitions from the neutral to a
infrared detector, non-dispersive infrared detector). Fig. 7 displays a specific ionic state. First results show a high sensitivity for radical
flow reactor used at Ghent University to study the pyrolysis of detection in combination with highly reliable isomer identifica-
alcohol fuels using several offline GCs. tion even for large species and multicomponent mixtures [120].
More recently, molecular-beam sampling into a differentially- Table 3 presents a comprehensive summary of fundamental
pumped mass spectrometer (MBMS) [113e117] has proven a combustion experiments conducted on C1 to C6 alcohols, along
powerful tool for measuring a large number of stable and radical with the diagnostic techniques employed in each, and the experi-
species from alcohol combustion with minimal disturbance to the mental conditions (e.g., temperature, pressure, and mixture

Fig. 7. Schematic of the experimental pyrolysis setup at Ghent University used to study pyrolysis of alcohols. 1: butanol vessel, 2: water vessel, 3: electronic balance, 4: pump, 5:
valve, 6: evaporator, 7: mixer, 8: heater, 9: air cylinder, 10: pressure regulator, 11: mass flow controller, 12: nitrogen cylinder, 13: reactor, 14: nitrogen internal standard, 15: oven, 16:
GC for formaldehyde and water, 17: GC for C5 and higher, 18: cyclone, 19: condenser, 20: dehydrator, 21: GC for C4 and lower, and 22: data acquisition.
Reprinted from Harper et al. [121] with permission from Elsevier.
50 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

presence of carbonyl products (e.g., ketones and aldehydes), eluci-


dating reaction mechanisms, and deriving kinetic rate coefficients.
These pioneering studies were motivated by the chemist’s fasci-
nation with various organic molecule reactions; there was no
mention of the relevance of such work to alcohol as a fuel for en-
ergy applications. Nevertheless, these early experimental works
contributed to laying foundations for understanding the combus-
tion chemistry of alcohols, including the development of experi-
mental methods, diagnostic techniques, reaction mechanisms, and
rate coefficient evaluations. Together with theoretical work by
Hinshelwood et al. [258] on chemical kinetics theory and by
Semenov [259] on chain reactions, a detailed understanding of the
combustion chemistry of alcohols was made possible. Indeed, the
water elimination pathways observed by Schultz and Kistiakowsky
[257] were recently shown to affect the ignition delay times at high
temperatures of tert-butanol [260], and the greater concentrations
of carbonyl products seen by Barnard [189,201,202] during alcohol
pyrolysis when compared with alkane pyrolysis explains why these
Fig. 8. Number of refereed journal articles published per year on various alcohol fuels. emissions are a source of concern in alcohol-fueled engines.
Alcohol isomers and alcohols larger than C5 are lumped together in the figure.
Articles published in the 1970s through the 1990s reported re-
sults from alcohol combustion in experimental systems similar to
those used in laboratories today (e.g., shock tubes, flow reactors,
fraction). The remaining sub-sections in this section critically re- and premixed flames). The early works state, as their motivations,
view the large body of experimental data on alcohol fuels presented the need to understand the fundamental combustion chemistry of
in Table 3. The objectives of this review are as follows. alcohols because of their potential use as alternative sources of
fuels for internal combustion engines. The idea was then proposed
 To elucidate historical trends in on the combustion chemistry of that fundamental combustion experiments, on such topics as the
alcohol fuels. This is achieved by presenting a graph on the ignition delay time and the premixed laminar flame speed, could
number of papers published for each family of alcohols (e.g., aid practical engine design. These experimental studies were
methanol, ethanol, propanols, etc.) per year since 1970. largely limited to methanol and ethanol, and there was less than
 To identify the variety of experiments and the range of tem- one article published per year over the 30-year period from 1970 to
peratures, pressures, and mixture fractions studied for each of 2000.
the alcohols, with an aim to highlight gaps on which future work In 2003, there was a sharp increase in the number of articles
could focus. published per year on these topics. From 2003 to 2013, the number
 To identify exemplary experimental data sets that will aid in of articles published is more than double that of all prior years
testing the detailed chemical models for elucidating the com- combined. Ethanol became a topic of interest, while methanol and
bustion chemistry of alcohols. propanol isomers studies averaged of one or two papers per year.
 To reveal peculiar and intriguing experimental phenomena that The significant growth in interest was in butanol isomers, on which
require further investigation. there have been numerous papers published since 2008. Histori-
cally, there has been a scarcity of fundamental combustion data on
higher alcohols (i.e., C5 and larger), as is also true for high molecular
weight hydrocarbons (e.g., alkanes, aromatics, etc.). These low va-
3.2. Developments in experimental combustion research on por pressure compounds complicate fuel vaporization and subse-
alcohols quent condensation in transfer lines, thus making it challenging to
achieve well-defined mixture compositions in the reactor/flame.
Fig. 8 presents a graph on the number of refereed journal articles Experimentalists are overcoming these issues by using atomizers/
published each year (since 1930) that describe alcohol combustion nebulizers to produce ultra-fine droplets/aerosols together with
experiments. Accompanying the graph are 4-dimensional plots (see various heating methods to promote vaporization and prevent
Fig. 9aee) representing the approximate range of temperatures, condensation (e.g., diluent gas pre-heaters, heated transfer lines,
pressures, equivalence ratios, and years at which alcohol combus- heating jackets, etc.). Thus, the body of experimental data on higher
tion has been studied. Only premixed oxidation studies are pre- alcohols is expected to grow considerably over the next decade. The
sented in the 4-D plots (i.e., pyrolysis and non-premixed flame recent interest in C4 and larger alcohols is a result of the industry’s
studies are excluded). The earliest articles on alcohol combustion interest in using alcohols with higher energy densities in trans-
actually date back to the 1930s when researchers [254e256] were portation applications, the design of novel synthetic biological
studying the “slow combustion” of methanol at low temperatures routes to produce higher molecular weight alcohols, and the
(400e750  C) and sub-atmospheric pressures. These were the first combustion scientist’s ability to generate high-quality fundamental
studies to show that methanol thermally decomposes in two stages combustion data rapidly.
(CH3OH / H2 þ CH2O / 2H2 þ CO), as well as to measure the re-
action rates. Around the same time, Schultz and Kistiakowsky [257] 3.3. Methanol, ethanol, and isomers of propanol
studied the thermal decomposition of tert-butanol and tert-iso-
amyl alcohol (1,1-dimethylpropan-1-ol) at 487e555  C and sub- The summary of fundamental combustion experiments given in
atmospheric pressures, identifying the low-lying molecular elimi- Table 3 provides an overview of the available studies relevant for
nation pathways forming an alkene and water. From 1957 to 1960, development and testing of gas-phase reaction kinetics for the
Barnard [189,201,202] continued to study the pyrolysis of alcohols respective alcohol fuel and its derived intermediates. The list in-
(iso-propanol, n-butanol, and tert-butanol), identifying the cludes data acquired for pure, gaseous combustion of the respective
Table 3
Fundamental experimental studies of alcohol combustion kinetics.

Reference Fuels Experimental device Measured properties Diagnostic techniques T, P, mixture fraction

Methanol
Cooke et al., 1971 [122] Methanol, ethanol Shock tube Ignition delay Emission of OH, CH, C2, and CO2 1570e1870 K,
0.267e0.400 bar, f ¼ 1
Bowman, 1975 [123] Methanol Shock tube Species profiles O, OH, H2O, and CO emission 1545e2180 K,
and absorption f ¼ 0.75e6.0
Aronowitz et al., 1979 [124] Methanol Flow reactor Species profiles GC 950e1030 K, 1 bar,
f ¼ 0.03e3.16
Singh et al., 1979 [125] Methanol Jet-stirred reactor Species profiles (f) GC (FID þ TCD), commercial 1870 K, 0.93 bar f ¼ 0.7e1.4
analyzers
Natarajan and Bhaskaran, Methanol Shock tube Ignition delay Wideband emission 1300e1700 K,
1981 [126] 2.5e4.5 bar, f ¼ 0.5e1.5
Vandooren and van Methanol Laminar premixed flame Species profiles (x) EI-MBMS, 500e2100 K, 0.053 bar,
Tiggelen, 1981 [127] f ¼ 0.89, 0.36, and 0.21

S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102
Tsuboi and Hashimoto, Methanol Shock tube Ignition delay and Emission between 2.7 and 5.5 mm w1200e1800 K, f ¼ 0.2e2.0
1981 [128] species profiles CH3OH, H2O, CO2, CO, CH2O
Gülder, 1982 [129] Methanol, ethanol Constant volume bomb Laminar burning velocities Ionization probes w300e600 K,
iso-octane 1e8 bar, f ¼ 0.7e1.4
Metghalchi and Keck, Methanol, iso-octane, Constant volume bomb Laminar burning velocities Ionization probes 298e700 K,
1982 [130] indolene 0.4e50 bar, f ¼ 0.8e1.5
Pauwels et al., 1989 [131] Methanol Premixed flame Species profiles (x) GC and EPR 0.107 bar, f ¼ 1.0
Norton and Dryer, Methanol Turbulent flow reactor Species profiles (x, t) GC, infrared and thermo- w1000e1150 K,
1989 [132] magnetic detection 1 bar, f ¼ 0.6e1.6
Norton and Dryer, Methanol, ethanol, n-, Flow reactor Species profiles (x, t) GC 1020e1120 K (initial),
1991 [133] iso-propanol, 1 bar, f ¼ 1.18
tert-butanol,
MTBE
Bradley et al., 1991 [134] Methanol Premixed flames Burning velocities and GC and LDV (CH3OH, w400e2000 K,
species profiles O2, H2, CO2, and CO) 0.09e0.253 bar,
f ¼ 0.7e1.3
Cribb et al., 1992 [135] Methanol Shock tube pyrolysis Refractive index gradient LS and MS (CH3OH, CH2O, 1800e2800 K,
of incident shock and CO, C2H2, CH4, CH3, and H2O) 0.28e0.86 bar
species profiles of the
reflected shock
Cribb et al., 1992 [136] Methanol Shock tube Refractive index gradient LS and MS (CH3OH, CH2O, 1800e2800 K,
of incident shock and CO, C2H2, CH4, CH3, and H2O) 0.655e0.873 bar,
species profiles of the f ¼ 0.5e3
reflected shock
Egolfopoulos et al., 1992 [137] Methanol Counterflow premixed flame Laminar flame speed LDV Flame temperature,
1 bar, f ¼ 0.5e2.0
Lee et al., 1993 [138] Methanol, ethanol, Rapid compression machine Ignition delay Pressure 750e1000 K,
MTBE 20e40 bar, f ¼ 1
Aniolek and Wilk, 1995 [139] Methanol Constant volume reactor P,T, species time histories GC (FID þ TCD), FTIR 667e705 K, 1 bar,
(up to 900 s) f ¼ 0.5e1.5
Held and Dryer 1994 [140] Methanol Variable pressure flow reactor Species profiles Various spectroscopy 752e1043 K, 1e20 bar,
techniques f ¼ 0.4e2.6
Li and Williams, 1996 [141] Methanol Counterflow diffusion flame Species profiles (x), Microprobe-GC, PDPA 300e210 K, 1 bar,
velocity field strain rate 50/se100/s
Fieweger et al., 1997 [142] Methanol, iso-octane, MTBE Shock tube Ignition delay Pressure transducers w750e1250 K,
and CH-emission at 431 nm 13e40 bar, f ¼ 1.0
Alzueta et al., 2001 [143] Methanol, NO Flow reactor Species profiles (T) FTIR/NO analyzer 700e1500 K,
1 bar, f ¼ 0.1e2.70
Saeed and Stone, 2004 [144] Methanol Constant volume bomb Laminar burning velocities High-speed CCD 293e425 K,
0.5e3.5 bar,
f ¼ 0.8e1.6
Liao et al., 2006 [145] Methanol Constant volume bomb Laminar burning velocities High-speed CCD
(continued on next page)

51
Table 3 (continued )

52
Reference Fuels Experimental device Measured properties Diagnostic techniques T, P, mixture fraction

358e480 K; 1 bar,
f ¼ 0.7e1.4
Dayma et al., 2007 [146] Methanol, NO Jet-stirred reactor Species profiles (T) GC(FID) 700e1000 K, 10 bar,
f ¼ 0.3e1
Rasmussen et al., 2008 [147] Methanol Flow reactor Species profiles (x, t) UV and IR 650e1350 K, 1 bar,
spectrophotometry, O2 1e16%
Zhang et al., 2008 [148] Methanol Constant volume bomb Laminar burning High-speed schlieren 373e473 K, 1e7.5 bar,
velocities photography f ¼ 0.8e1.2
Togbé et al., 2009 [149] Methanol, Jet-stirred reactor Species profiles (T) GC (FID), FTIR 770e1140 K, 10 bar,
M85 surrogate f ¼ 0.35e2
Noorani et al., 2010 [150] Methanol, ethanol, Shock tube Ignition delays CH-emission 1070e1760 K,
n-propanol, 2e12 bar,
n-butanol f ¼ 0.75 and 1.8
Veloo et al., 2010 [110] n-Butanol, methanol, Counterflow premixed flame Laminar flame speed, PIV 343 K, 1 bar,

S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102
ethanol extinction strain rates f ¼ 0.7e1.5
Kumar and Sung, 2011 [151] Methanol Rapid compression machine Ignition delay times Pressure 7e30 bar, 850e1100 K,
f ¼ 0.25e2.0
Vancoillie et al., 2012 [152] Methanol Flat-flame burner Laminar flame velocities Heat-flux method Flame temperature,
1 bar, f ¼ 0.7e1.5
Aranda et al., 2013 [153] Methanol Flow reactor Species profiles (x, t) GC (FID and TCD) 600e900 K,
20e100 bar,
f ¼ 0.0625e4.0
Beeckmann et al., 2014 [154] Methanol, ethanol, Constant volume bomb Laminar flame speed High-speed schlieren 373 K, 10 bar,
n-propanol, photography f ¼ 0.7e1.3
n-butanol

Ethanol
Cooke et al., 1971 [122] Ethanol, methanol Shock tube Ignition delay Emission of OH, CH, 1570e1870 K,
C2 and CO2 0.267e0.400 bar,
f¼1
Gülder, 1982 [129] Methanol, ethanol, Constant volume bomb Laminar burning Ionization probes w300e600 K,
iso-octane velocities 1e8 bar, f ¼ 0.7e1.4
Rotzoll, 1985 [155] Ethanol Pyrolysis flow reactor Species profiles (T) MS (C2H5OH, C2H4, w1000e1300K 1.3 bar
CH3CHO,
H2, CH4)
Norton and Dryer, 1991 [133] Methanol, ethanol, n-, Flow reactor Species profiles (x, t) GC 1020e1120 K (initial),
iso-propanol, 1 bar, f ¼ 1.18
tert-butanol, MTBE
Dunphy and Simmie, 1991 [156] Ethanol Shock tube Ignition delay Pressure sensor 1060e1660 K, 1.8e4.6 bar,
and emission f ¼ 0.25e2
of CO þ O and OH
Curran et al., 1992 [157] Ethanol, iso-butene Shock tube Ignition delay Pressure transducer and 1100e1900 K,
and MTBE emission of OH or CO2 2e4.6 bar, f ¼ 0.25e1.5
Dagaut et al., 1992 [158] Ethanol Jet stirred reactor Species profiles (T) GC (CO, CO2, CH4, C2H4, w1000 K, 1 bar,
C2H6, CH3CHO, C2H5OH) f ¼ 0.2e2
Egolfopoulos et al., 1992 [159] Ethanol Counterflow premixed flame Laminar flame speed LDV Flame temperature,
1 bar, f ¼ 0.6e1.8
Norton and Dryer, 1992 [160] Ethanol Turbulent flow reactor Species profiles (x, t) GC w1100 K (initial),
1 bar, f ¼ 0.61e1.24
Lee et al., 1993 [138] Methanol, ethanol, MTBE Rapid compression machine Ignition delay Pressure 750e1000 K,
20e40 bar, f ¼ 1
Li et al., 2001 [161] Ethanol Flow reactor pyrolysis Species profiles (x, t) Online FTIR and 1045e1080 K,
offline GC 1.7e3.0 bar
Alzueta and Hernández, Ethanol, NO Flow reactor Species profiles (T) FTIR/NO Analyzer 700e1500 K,
2002 [162] 1 bar, f ¼ 0.1e2.70
Benvenutti et al., 2004 [163] Ethanol Low-pressure flame
Temporal species Light emission Flame temperature,
profiles of OH, HCO, 0.355 bar, f ¼ 1
CH2O, CH and C2
Hidaka et al., 2005 [164] Ethanol Shock tube pyrolysis Species profiles GC þ IR absorption 1000e1700 K,
1.4e3.6 bar
Ergut et al., 2006 [165] Ethanol, ethylbenzene Premixed flame Species profiles, PAH/soot absorption; w1800 K, 1 bar, f ¼ 2.5
PAH, soot (x) GC (MS, FID, TCD),
Kasper et al., 2007 [166] Ethanol Low-pressure premixed flame Species profiles (x) EI- and REMPI eMBMS; w400e2400 K, 0.05 bar,
f ¼ 1.00 and 2.57
Liao et al., 2007 [167] Ethanol Constant volume bomb Laminar burning High-speed CCD 358e480 K; 1e10 bar,
velocities f ¼ 0.7e1.4
Saxena and Williams, 2007 [168] Ethanol Partially premixed and Species profiles (x) GC w300e2000 K, 1 bar
non-premixed counterflow flame
Dagaut and Togbé, 2008 [169] Ethanol, E85 surrogate Jet-stirred reactor Species profiles (T) GC (FID), FTIR 800e1100 K, 10 bar,
f ¼ 0.6e2
Leplat et al., 2008 [170] Ethanol Low-pressure premixed flame Species profiles (x) EI-MBMS w600e2000 K,
0.05 bar, f ¼ 1.00

S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102
Wang et al., 2008 [171] Ethanol, DME, propene Laminar premixed flame Species profiles (x) MBMS (PI, EI), w500e2200 K,
0.04 bar, f ¼ 2
Bradley et al., 2009 [172] Ethanol Constant volume bomb Laminar burning High-speed schlieren 300e393 K, 1e14 bar,
velocities photography f ¼ 0.7e1.5
Cancino et al., 2010 [173] Ethanol Shock tube Ignition delays CH-emission 650e1220 K, 10, 30, and
50 bar, f ¼ 0.3
Noorani et al., 2010 [150] Methanol, ethanol, Shock tube Ignition delays CH-emission 1070e1760 K, 2e12 bar,
n-propanol, f ¼ 0.75 and 1.8
n-butanol
Veloo et al., 2010 [110] n-Butanol, methanol, Counterflow premixed flame Laminar flame speed, PIV 343 K, 1 bar, f ¼ 0.7e1.5
ethanol extinction strain rates
Heufer and Olivier, 2010 [174] Ethanol Shock tube Ignition delay Pressure Down to 800 K,
13e40 bar, f ¼ 1.0
Eisazadeh-Far et al., 2011 [175] Ethanol Cylindrical and spherical vessel Laminar burning speed Shadowgraph system 300e650 K, 1e5 bar,
with CMOS camera f ¼ 0.8e1.1
Konnov et al., 2011 [176] Ethanol Flat-flame burner Laminar flame velocities Heat-flux method Flame temperature,
1 bar, f ¼ 0.7e1.5
Leplat et al., 2011 [177] Ethanol Low-pressure premixed Species profiles (x) EI-MBMS, GC w500e1800 K, 0.05 bar,
flame and jet-stirred reactor f ¼ 0.75, 1, and 1.25;
890e1250 K, 1 bar,
f ¼ 0.25, 0.5, 1, and 2
van Lipzig et al., 2011 [178] Ethanol, n-heptane, Flat-flame burner Laminar burning velocities Heat-flux 298e338 K, 1 bar,
iso-octane (thermocouples) f ¼ 0.6e1.5
Xu et al., 2011 [179] Ethanol, DME Low-pressure premixed flame Species profiles (x) PI-MBMS w500e2000 K, 0.04 bar,
f ¼ 1.00
Broustail et al., 2011 [180] Ethanol, butanol Constant volume bomb Laminar burning velocities High-speed 393 K, 1 bar,
iso-octane Shadowgraph f ¼ 0.8e1.4
Lee et al., 2012 [181] Ethanol Shock tube and rapid Ignition delays Pressure, CH-emission, 775e1300 K, w80 bar
compression machine schlieren
Varea et al., 2012 [182] Ethanol Constant volume reactor Laminar flame speed High-speed laser 373 K, 1e5 bar,
tomography f ¼ 0.7e1.5
Broustail et al., 2013 [183] Ethanol, butanol iso-octane Constant volume bomb Laminar burning velocities High-speed 423 K, 10 bar,
Shadowgraph f ¼ 0.8e1.4
Tran et al., 2013 [184] Ethanol, methane Low-pressure flame Species profiles (x) Microprobe-GC w300e2100 K,
(FID þ TCD þ MS) 0.066 bar, f ¼ 0.7e1.3
Varea et al., 2013 [185] Ethanol, iso-octane blends Constant volume reactor Laminar flame speed Mie scattering 373 K, 1e10 bar,
f ¼ 0.7e1.5
Beeckmann et al., 2014 [154] Methanol, ethanol, Constant volume bomb Laminar flame speed High-speed schlieren 373 K, 10 bar,
n-propanol, n-butanol photography f ¼ 0.7e1.3
Herrmann et al., 2014 [186] Ethanol, DME Flow reactor Species profiles (T) TOF-MS (EI) 400e1200 K, 1 bar,
f ¼ 0.8e1.2
Mittal et al., 2014 [187] Ethanol Rapid compression machine Ignition delay times Pressure 10e50 bar, 825e985 K,
f ¼ 0.3e1.0

53
(continued on next page)
Table 3 (continued )

54
Reference Fuels Experimental device Measured properties Diagnostic techniques T, P, mixture fraction

Knorsch et al., 2014 [188] Ethanol, n-, iso-butanol, Heat flux burner Laminar flame speed Heat-flux method 373e423 K, 1 bar,
iso-octane, n-heptane f ¼ 0.5e1.5
Propanol isomers
Barnard, 1960 [189] iso-Propanol Batch reactor pyrolysis Decomposition rate, Pressure, classical 797e888 K, 0.027e0.400 bar
products chemistry, GC
Barnard and Hughes, 1960 [190] n-Propanol Batch reactor pyrolysis Decomposition rate, Pressure, classical 843e895 K, 0.020e0.400 bar
products chemistry, GC
Norton and Dryer, 1991 [133] Methanol, ethanol, n-, Flow reactor Species profiles (x, t) GC 1020e1120 K (initial),
iso-propanol, tert-butanol, 1 bar, f ¼ 1.18
MTBE
Sinha and Thomson, 2004 [191] iso-Propanol, dimethoxy Counterflow diffusion flame Species profiles (x) Microprobe-GC(FID), 400e1800 K, 1 bar
methane, HPLC (UV), NDIR-sensor;
dimethyl carbonate
Li et al., 2008 [192] n- and iso-Propanol, acetone Laminar premixed flame Species profiles (x) MBMS(PI, EI) 500e2000 K, 0.02 and

S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102
0.04 bar, f ¼ 0.75 and 1.8
Johnson et al., 2009 [193] n- and iso-Propanol Shock tube Ignition delay Pressure, CH-emission 1350e2000 K, 1 bar,
f ¼ 0.5e2
Kasper et al., 2009 [194] n- and iso-Propanol Laminar premixed flame Species profiles (x) MBMS(PI, EI) 500e2000 K,
0.033e0.047 bar,
f ¼ 1.0e1.9
Frassoldati et al., 2010 [195] n-Propanol, iso-propanol Counterflow diffusion flame Species profiles (x) Microprobe, GC(TDC) Flame temperature,
1 bar, strain rate 97.5 1/s
Noorani et al., 2010 [150] n-Propanol, methanol, Shock tube Ignition delays CH-emission 1070e1760 K, 2e12 bar,
ethanol, f ¼ 0.75 and 1.8
n-butanol
Akih-Kumgeh and Bergthorson, iso-Propanol, propanal, Shock tube Ignition delay CH-emission 1250e1700 K, 1 and
2011 [196] acetone, ethyl formate 12 bar, f ¼ 0.5e2
Galmiche et al., 2011 [197] n-Propanol Jet-stirred and batch reactor Species profiles (T) and FTIR, GC(FID, TCD, MS) 770e1190 K, 10 bar,
laminar flame speed f ¼ 0.35e2; Flame
speed: 323e473 K,
1e10 bar, f ¼ 0.7e1.4
Veloo et al., 2011 [198] n-Propanol, iso-propanol, Counterflow premixed flame Laminar flame speed Symmetric twin-flame 343 K, 1 bar,
propane f ¼ 0.7e1.5
Esarte et al., 2012 [199] Methanol, ethanol, Flow reactor pyrolysis and Species profiles (T) GC(FID þ TCD) 975e1475 K,
iso-propanol, oxidation with acetylene 1 bar, f ¼ 1
n-butanol
Man et al., 2014 [200] n-Propanol and Shock tube Ignition delay OH-emission 1100e1500 K,
iso-propanol 1e26 bar, f ¼ 0.5e2
Beeckmann et al., 2014 [154] Methanol, ethanol, Constant volume bomb Laminar flame speed High-speed schlieren 373 K, 10 bar,
n-propanol, photography f ¼ 0.7e1.3
n-butanol

Butanol isomers
Barnard, 1957 [201] n-Butanol Batch reactor pyrolysis Decomposition rate, Pressure, classical 846e902 K,
products chemistry, GC 0.027e0.667 bar
Barnard, 1959 [202] tert-Butanol Batch reactor pyrolysis Decomposition rate, Pressure, classical 760e893 K,
products chemistry, GC 0.027e0.533 bar
Tsang, 1964 [203] tert-Butanol, tert-butyl-chlonde, Single-pulse shock tube Decomposition rate GC 1050e1300 K,
-bromide, and -mercaptan pyrolysis 0.533e1.333 bar
Lewis et al., 1974 [204] tert-Butanol Single-pulse shock tube Decomposition rate GC 920e1175 K,
pyrolysis 0.5e2.1 bar
Choudhury et al., 1990 [205] tert-Butanol Shock tube pyrolysis Time histories (CO Absorption 1220e1620 K,
and H2O) spectroscopy 0.4e1.1 bar
Norton and Dryer, 1991 [133] Methanol, ethanol, n-, Flow reactor Species profiles (x, t) GC 1020e1120 K (initial),
iso-propanol, tert-butanol, 1 bar, f ¼ 1.18
MTBE
McEnally and Pfefferle, 2005 [206] n-,2-, iso- and tert-Butanol, Coflow methane flame Centerline species Microprobe-MS (EI, 1 bar
n-butane, iso-butane profiles (x) 118 nm PI)
Moss et al., 2008 [207] n-,2-, iso- and tert-Butanol Shock tube Ignition delay Pressure, OH-emission 1200e1800 K,
1e4 bar, f ¼ 0.25e1
Dagaut et al., 2009 [208] n-Butanol Jet-stirred reactor Species profiles (T) GC (FID, TCD), FTIR 750e1100 K,
10 bar, f ¼ 0.5e2
Dagaut and Togbé, 2009 [209] n-Butanol/n-heptane mixture Jet-stirred reactor Species profiles (T) GC (FID), FTIR 530e1070 K,
10 bar, f ¼ 0.5 and 1
Gu et al., 2009 [210] n-Butanol Batch reactor Flame speed High-speed schlieren 413- 473 K, 1.0 bar
photography and 2.5 bar,
f ¼ 0.8e1.6
Sarathy et al., 2009 [211] n-Butanol Jet-stirred reactor, Species profiles (T and x), FTIR, GC (FID, TCD), 400e2000 K
counterflow diffusion flame speed Microprobe-GC (Flame),
flame, batch reactor (FID), NDIR, HPLC (UVeVis) 850e1250 K (JSR),
1 bar, f ¼ 0.25e2
Black et al., 2010 [212] n-Butanol Shock tube Ignition delay time Pressure, emission 1100e1800 K,
1e8 bar, f ¼ 0.5e2
Grana et al., 2010 [213] n-,2-, iso- and tert-Butanol Counterflow diffusion flame Species profiles (x) Microprobe-GC (TCD) 400e1800 K,
1 bar, strain rate 100/s

S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102
Gu et al., 2010 [214] n-,2-, iso- and tert-Butanol Batch reactor Flame speed High-speed schlieren 428 K, 1e7.5 bar,
photography f ¼ 0.7e1.5
Noorani et al., 2010 [150] n-Butanol, methanol, ethanol, Shock tube Ignition delays CH-emission 1070e1760 K,
n-propanol 2e12 bar,
f ¼ 0.75 and 1.8
Togbé et al., 2010 [215] 2- and iso-Butanol Jet-stirred reactor Species profiles (T) FTIR, GC (FID, TCD) 770e1250 K,
10 bar, f ¼ 0.275e4
Van Geem et al., 2010 [216] n-,2- and tert-Butanol Flow reactor pyrolysis Species profiles (x, T) GC x GC (FID, TCD, MS) 880e1010 K,
1.5e1.7 bar
Veloo et al., 2010 [110] n-Butanol, methanol, ethanol Counterflow premixed flame Laminar flame speed, PIV 343 K, 1 bar,
extinction f ¼ 0.7e1.5
strain rates
Broustail et al., 2011 [180] n-Butanol, ethanol iso-octane Constant volume bomb Laminar burning velocities High-speed Shadowgraph 393 K, 1 bar,
f ¼ 0.8e1.4
Gu et al., 2011 [217] tert-Butanol Batch reactor Flame speed High-speed schlieren 428e488 K,
photography 1e5 bar, f ¼ 0.8e1.5
Hansen et al., 2011 [218] n-Butanol, n-butanol/H2 Laminar premixed flame Species profiles (x) MBMS (PI) w500e1900 K,
0.020e0.033 bar,
f ¼ 1.0e1.4
Harper et al., 2011 [121] n-Butanol Flow reactor pyrolysis Species profiles (T) GC x GC (FID, TCD, MS) 600e807 K,
1.7 bar
Heufer et al., 2011 [219] n-Butanol Shock tube Ignition delay Pressure, CH-emission 770e1250 K,
10e42 bar, f ¼ 1
Karwat et al., 2011 [220] n-Butanol Rapid compression machine Ignition delay, product Pressure, luminescence 920e1040 K,
speciation (reconstructed imaging, GC (FID, TCD) 2.90e3.39 bar
time histories)
Obwald et al., 2011 [221] n-,2-, iso- and tert-Butanol Laminar premixed flame Species profiles (x) MBMS (PI, EI) 500e2000 K,
0.040 bar, f ¼ 1.7
Veloo and Egolfopoulos, n-,2-, iso- and tert-Butanol Counterflow premixed flame Laminar flame speed, PIV 343 K, 1 bar,
2011 [222] extinction strain rates f ¼ 0.7e1.5
Vranckx et al., 2011 [223] n-Butanol Shock tube Ignition delay Pressure, CH-emission 795e1200 K,
61e92 bar, f ¼ 1
Weber et al., 2011 [104] n-Butanol Rapid compression machine Ignition delay Pressure 675e925 K,
15e30 bar,
f ¼ 0.5e2
Cai et al., 2012 [224] tert-Butanol Flow reactor pyrolysis Species profiles (T) PI-MBMS 950e1850 K,
0.004e0.016 bar
Cai et al., 2012 [225] n-Butanol Flow reactor pyrolysis and Species profiles (T, x) PI-MBMS 800e1500 K,
laminar premixed flame 0.007e1.013 bar;
Flame: 0.040 bar,
f ¼ 0.7e1.8
Esarte et al., 2012 [199] Species profiles (T) GC (FID þ TCD)
(continued on next page)

55
Table 3 (continued )

56
Reference Fuels Experimental device Measured properties Diagnostic techniques T, P, mixture fraction

Methanol, ethanol, Flow reactor pyrolysis and 975e1475 K,


iso-propanol, n-butanol oxidation with acetylene 1 bar, f ¼ 1
Karwat et al., 2012 [226] n-Butanol/n-heptane blends Rapid compression machine Ignition delay, product Pressure, luminescence 700 K, 9 bar, f ¼ 1
speciation imaging, GC (FID, TCD)
(reconstructed time
histories)
Lefkowitz et al., 2012 [227] tert-Butanol Flow reactor and counterflow Species profiles (T, x), FTIR, NDIR, GC (FID) 675e950 K, 12
diffusion flame extinction and 5 bar f ¼ 1;
strain rate Flame: 1 bar, strain
rate 100/s
Liu et al., 2011 [228] n-Butanol, iso-butanol, Batch reactor and liquid Laminar flame speed, High-speed schlieren 353 K, 1 and 2 bar, f ¼ 0.7e1.4
methyl butanoate pool stagnation burner ignition photography
temperature
Stranic et al., 2012 [108] n-,2-, iso- and tert-Butanol Shock tube Ignition delay Pressure, OH-emission 1050e1600 K,

S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102
1.0e44 bar,
f ¼ 1 and 0.5
Stranic et al., 2012 [111] n-Butanol Shock tube pyrolysis Multi-species time Laser absorption (OH, H2O, 1250e1650 K,
histories C2H4, CO, and CH4) 1.3e1.9 bar
Van Geem et al., 2012 [229] n-Butanol, acetone/ethanol/ Flow reactor pyrolysis Species profiles GC x GC (FID, TCD, MS) 880e1010 K,
butanol-mixture (conversion, T) 1.5e1.7 bar
Yasunaga et al., 2012 [230] n-,2-, iso- and tert-Butanol Single-pulse shock tube Speciation (T) GC (TCD) 1000e1800 K,
1e4 bar, f ¼ 0.25e1
Broustail et al., 2013 [183] Ethanol, n-butanol iso-octane Constant volume bomb Laminar burning velocities High-speed Shadowgraph 423 K, 10 bar,
f ¼ 0.8e1.4
Chung et al., 2013 [231] n-, iso-Butanol and n-, iso-butene Premixed stagnation flame NO, velocity field NO-LIF, PIV Flame temperature,
1 bar, f ¼ 0.8e1.3
Hansen et al., 2013 [232] iso-Butanol Laminar premixed flame Species profiles (x) MBMS (PI) w500e1900 K,
0.015e0.03 bar,
f ¼ 1.0e1.5
Merchant et al., 2013 [233] iso-Butanol Flow reactor pyrolysis Species profiles (T) GC x GC (FID, TCD, MS) 550e1093 K,
1.7 bar
Zhang et al., 2013 [234] n-Heptane/n-butanol mixtures Shock tube Ignition delay Pressure, CH-emission 1200e1500 K,
2e10 bar, f ¼ 0.5 and 1
Stranic et al., 2013 [112] iso- and 2-Butanol Shock tube pyrolysis Multi-species time histories Laser absorption (OH, H2O, 1270e1640 K,
C2H4, CO, and CH4) 1.3e1.9 bar
Cai et al., 2013 [235] 2-Butanol Flow reactor pyrolysis and Species profiles (T, x) PI-MBMS 800e1500 K,
laminar premixed flame 0.007e1 bar; Flame:
0.04 bar, f ¼ 0.7e1.8
Wu and Law, 2013 [236] n-, 2-, iso- and tert-Butanol Constant volume bomb Laminar flame speed High-speed schlieren 353 and 373 K,
photography 1e5 bar, f ¼ 0.8e1.4
Camacho et al., 2013 [237] n-, iso-Butanol and n-, iso-butene Premixed stagnation flame Particle size distribution (x) SMPS 400e1800 K,
1.0 bar, f ¼ 2.25
Vasu and Sarathy, 2013 [238] n-Butanol Shock tube OH time histories Laser absorption of OH 1300e1550 K,
2 bar, f ¼ 1
Yang et al., 2013 [239] n-Heptane/n-butanol mixtures Rapid compression machine Ignition delay Pressure 650830 K,
15e30 bar,
f ¼ 0.4e1.5
Zhu et al., 2014 [240] n-Butanol Shock tube Ignition delay Pressure, OH-emission, 716e1121 K,
IR-absorption 0.020 and 0.040 bar,
f ¼ 0.5e2
Knorsch et al., 2014 [188] Ethanol, n-, iso-butanol Heat flux burner Laminar flame speed Heat-flux method 373e423 K,
iso-octane, n-heptane 1 bar, f ¼ 0.5e1.5
Cai et al., 2014 [241] iso-Butanol Flow reactor pyrolysis Species profiles (T) PI-MBMS 950e1850 K,
0.004e0.016 bar
Jin et al., 2013 [242] n-,2-, iso- and tert-Butanol Coflow diffusion flame Centerline species profiles (x) PI-Microprobe-MS w500e1700 K, 1 bar
Pan et al., 2014 [243] iso-Butanol Shock tube Ignition delay Pressure, CH-emission 900e1700 K,
1.2e10 bar,
f ¼ 0.5e1
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 57

alcohol under well-defined conditions and provides properties


such as ignition delay times, laminar flame speeds, and species
concentration profiles. Beyond these kinetics-related in-

1.01 bar, strain rate 33/s


770e1220 K and 423 K,
vestigations, a substantial number of investigations focusing on the

1e2.6 bar, f ¼ 0.25e1

5e60 bar, f ¼ 0.35e4


530e1220 K, 10 bar,

10 bar, f ¼ 0.7e1.4
combustion properties of the liquid fuel (e.g. [261],) or the influence

10 bar, f ¼ 0.5e2

473 K, 1e7.5 bar,


600e1100 K and

of alcohol addition to hydrocarbon fuels (e.g. [262e266],) are not

9e30 bar, f ¼ 1

1100e1500 K,
7e23 bar, and
373 K, 10 bar,

423 K, 10 bar,

393, 433, and


considered herein. Similarly, studies on the “slow combustion” of
560e1030 K,

640e1200 K,

652e1457 K,

400e1800 K,
530e1800 K,
f ¼ 0.7e1.3

f ¼ 0.5e3.5

f ¼ 0.5e2.0

f ¼ 0.6e1.8
f ¼ 0.35e4

alcohols, meaning a reaction time of several seconds up to some


minutes conducted in the 1950’s (e.g. [267]), and studies on
oxidation of alcohols in supercritical water [268e271] are excluded
herein.
Cathonnet et al. [272,273] conducted a series of experimental
studies on the oxidation and pyrolysis of methanol. Ignition delay
time measurements are available for temperatures ranging from

Microprobe-GC (FID, TCD),


700 K to 2800 K, a stoichiometry range of f ¼ 0.2e6, and pressures

FTIR, GC (FID, TCD, MS),


GC (TCD, FID, MS) FTIR,

FTIR, GC (FID, TCD, MS)

GC (TCD, FID, MS) FTIR,

FTIR, GC (FID, TCD, MS)


High-speed schlieren

High-speed schlieren
up to 40 bar are reported. The majority of ignition delay data was

temperature (TC)
acquired from shock tubes in the 1970s and early 1980s. Extensive
Shadowgraphy

Shadowgraphy

PIV, pressure studies on shock tube ignition of methanol are provided by


photography

photography
CH-emission

Bowman [123] and Tsuboi et al. [128]. Both cover a temperature


Pressure

Pressure

range between 1500 and 2000 K and a wide stoichiometry range


and offer parameterized expressions of their results. Bowman [123]
additionally provides concentration time histories for some species.
More recent studies are scarce; RCM results [151] were obtained in
Species profiles (T), ignition
Species profiles (T), laminar

Species profiles (T), laminar

delay, laminar flame speed,

the low-temperature regime, while other studies focus on a wide


range of topics and provide just a limited number of conditions for
Laminar burning speed

extinction strain rate


Laminar flame speed

methanol [142,150,274].
Species profiles (T)

Species profiles (T)

Species profiles (x)

Ignition delay times for ethanol are available up to 1900 K and


Ignition delay

Ignition delay

Ignition delay

80 bar in the stoichiometry range of f ¼ 0.25e2. Studies in the


flame speed

flame speed

high-temperature regime date back the 1990s and earlier spanning


twenty-nine individual conditions, Dunphy and Simmie [275]
provide the largest dataset for ethanol ignition. The parameter-
Higher alcohols

ized ignition delay times are determined in a range of 1000e


1700 K, f ¼ 0.5e2.0, and pressures between 2 and 4.6 bar. More
Jet-stirred reactor, shock tube,

recent investigations (e.g. [173,181]) of ethanol ignition are focused


Counterflow diffusion flame
rapid compression machine

on the lower temperature region including results from RCMs.


Constant volume bomb

Overall, the amount of available ignition delay time measurements


compression machine

compression machine

is somewhat larger than for methanol.


Jet-stirred reactor,

Jet-stirred reactor,
Jet-stirred reactor

Jet-stirred reactor

Shock tube, rapid

Shock tube, rapid


cylindrical vessel

Propanol is the smallest alcohol existing in different isomeric


Spherical vessel
spherical vessel

structures (n- and iso-propanol). Nevertheless, the available exper-


Shock tube

imental data for both propanols are limited. Ignition delay time
measurements are just provided by four studies [150,193,196,200].
Johnson et al. [193] provide a direct comparison of both isomers, and
they report a higher reactivity of n-propanol compared to iso-prop-
anol. A comparative study of all normal alcohols (including meth-
Methanol, ethanol, n-propanol,

anol, ethanol and n-propanol) was recently provided by Noorani


n-Hexanol/jet A-1 mixtures

et al. [150] with ignition delay times measured under comparable


n-Pentanol, iso-pentanol,

conditions. They concluded that ignition delay times for the alcohols
n-Pentanol, n-hexanol

2-Methyl-1-butanol
2-methyl-1-butanol

(except for methanol) would be identical for a given stoichiometry


which, however, is not correctly reproduced by the kinetic models
tested in their study. A similar conclusion is drawn by Veloo et al.
iso-Pentanol

iso-Pentanol

iso-Pentanol
n-Pentanol
n-Hexanol

n-Hexanol
n-butanol

[110] when investigating laminar flame speeds in the lean-to-


stoichiometric regime of methanol, ethanol and n-butanol.
Laminar flame speeds for the C1eC3 alcohols are available for
the stoichiometry range of 0.7e1.5 and for pressures up to 10 bar.
Flame speed measurements are accessible by a broad variety of
experiments and are typically measured with air as the oxidizer. For
Mzé-Ahmed et al., 2012 [247]
Beeckmann et al., 2014 [154]

methanol and ethanol, several groups have measured experimental


Sarathy et al., 2013 [252]
Dayma et al., 2011 [245]

Heufer et al., 2012 [248]

data under comparable conditions. Recent results are available


Togbé et al., 2010 [244]

Togbé et al., 2011 [246]

Tang et al., 2013 [250]


Tsujimura, 2012 [249]

Yeung and Thomson,

using high-speed schlieren measurements in constant volume


Li et al., 2013 [251]

combustion chambers [145,148,167], from heat flux burner mea-


surements [152,176], and flow field determination by PIV (particle
2013 [253]

imaging velocimetry) in a counterflow premixed flame setup [110].


In general the database is a little larger for ethanol than for
methanol while for propanol just two papers report flame speeds
[197,198].
58 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Fig. 9. Four-dimensional plots representing the approximate range of temperatures, pressures, equivalence ratios, and years at which alcohol combustion has been studied. (a)
methanol, (b) ethanol, (c) propanol isomers, (d) butanol isomers, (e) higher alcohols. Only premixed oxidation studies are presented in the plots (i.e., pyrolysis and non-premixed
flame studies are excluded). The temperature of premixed flame studies are approximated as 2000 K.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 59

Quantitative species profiles are the most challenging test case quantitative data and cross-comparisons of important aldehydes,
for kinetic models. Speciation data of methanol combustion from a enols, and alkenes produced during the combustion of the four
variety of different flame and reactor experiments are available for butanol isomers. The premixed laminar flame speeds of the butanol
temperatures and pressures up to 2100 K and 100 bar. For meth- isomers have also been measured at a wide range of conditions
anol, the quantitative data are primarily available from reactor [110,150,210,214,217,228,236].
experiments for temperatures below 1000 K. Typically some or all Other detailed speciation studies include experiments in ideal
product and reactant species are traced, for example in the most reactors under oxidation and pyrolysis conditions. Jet-stirred
recent study of Aranda et al. [153] that extends the database up to [208,211,215] and flow reactor [227] oxidation data is available at a
pressures of 100 bar. Additionally, concentrations of formaldehyde range of pressures, temperatures, and equivalence ratios for all
and some other intermediate species (CH4, H2, CO) are determined isomers of butanol. Atmospheric pressure pyrolysis flow reactor
in most reactor studies. Compared to reactor experiments, quan- data has also been presented for all butanol isomers
titative data from higher temperature flame environments [216,224,225,233]. Species profiles during homogenous gas-phase
[127,131,134] are surprisingly scarce for methanol. The most ignition have been obtained in RCMs [220] and STs [111,112,230].
detailed investigation from Vandooren and van Tiggelen [127] The ignition delay times of the butanol isomers have also
dates back to 1981 and provides fourteen species profiles ob- received significant attention in the past 7 years. Moss et al. [207]
tained for three lean premixed flames. studied the high-temperature ignition characteristics of all four
Speciation data for ethanol is available for a wide range of butanol isomers in shock tubes, and concluded that tert-butanol is
conditions and experiments. Pressures up to 10 bar and a temper- the least reactive isomers while n-butanol is the most reactive.
ature range up to 2400 K have been covered. Several MBMS These high-temperature results were later confirmed and extended
[166,170,177,276e278] and other flame-sampling methods to wider ranges of pressures, temperatures, and equivalence ratios
[165,168,184] provide detailed insights to the high-temperature [108,212,230,240]. Although of relevance to engines, data for low-
regime. In contrast to methanol, higher molecular species (e.g., and intermediate-temperature ignition of the butanol isomers are
higher aldehydes such as acetaldehyde and soot precursors such as limited. Weber et al. [104,105] conducted RCMs studies of all
benzene) are detectable in rich ethanol flames. This is related to butanol isomers under high pressures and low temperatures of
ethanol reaction pathways forming C2 hydrocarbons such as direct relevance to reciprocating engines. Only n-butanol has been
ethylene or ethyl radicals. Lower temperatures are covered by a studied under such conditions in STs [219,223,240] because igni-
similar number of various reactor studies. Even though no typical tion delay times for the other butanol isomers exceed maximum
low-temperature regime exists for ethanol, recent results confirm a test times unless driver gas tailoring is employed.
distinct reactivity even at temperatures below 800 K [186,279],
which is rarely covered by most kinetic models. 3.5. Higher alcohols
Again for the propanol isomers, only a limited number of
speciation studies are available. Conditions cover pressures up to As mentioned earlier, the experimental data available on higher
10 bar and temperatures up to 2000 K, while no speciation data was alcohols (C5 and larger) are limited. For the pentanol isomers, there
reported below 750 K. Extensive studies are provided by Kasper are data on n-pentanol [103,246], 3-methyl-1-butanol
et al. [194] and Li et al. [192], both providing a large number of [245,249,250,252] (i.e., iso-pentanol), and 2-methyl-1-butanol
stable and radical species for flames of both isomers measured [250,251]. Togbé et al. [246] studied n-pentanol and Dayma et al.
using MBMS setups with isomer-specific detection. studied iso-pentanol combustion [245] in a jet-stirred reactor at
10 bar and a range of temperatures and equivalence ratios. n-
3.4. Butanol isomers Pentanol produced 1-butene and pentanal, whereas iso-pentanol
produced more iso-butene and 3-methyl-1-butanal. Similar to their
The four isomers of butanol are n-butanol, iso-butanol, 2- other studies on butanol isomers, n-pentanol and iso-pentanol did
butanol, and tert-butanol. Prior to the 1990s, literature studies on not show low-temperature reactivity under these experimental
the butanol isomers were limited to pyrolysis conditions. Inter- conditions. Their data on various primary alcohols (namely,
estingly, the primary focus was to determine decomposition rate of ethanol, n-propanol, n-butanol, iso-butanol, n-pentanol, iso-pen-
tert-butanol pyrolysis using batch reactors and shock tubes [202e tanol, and n-hexanol) revealed that they all produced similar in-
205]. Norton and Dryer [133] were the first to reveal species pro- termediate molar fractions of formaldehyde, whereas secondary
files evolving from the atmospheric-pressure oxidation of tert- alcohols (namely, iso-propanol and 2-butanol), yielded less form-
butanol in a flow reactor. aldehyde. Ethanol was shown to produce more acetaldehyde than
A growing interest in the combustion of butanol isomers began the other tested alcohols, regardless of their structure. Similar JSR
in the mid-2000s with McEnally and Pfefferle’s [206] study on the experiments on n-hexanol by the same group [244] yielded similar
decomposition and hydrocarbon growth processes of non- speciation results, except that hexanal was also observed. n-Hex-
premixed co-flow diffusion methane flames doped with anol exhibited low-temperature cool flame reactivity under the
3500 ppm of each of the four isomers of butanol. Gas samples tested conditions, which was not seen in smaller alcohols. Yeung
extracted from the flame revealed that the butanol isomers pro- et al. [253] also measured species generated during combustion of
duce much higher concentrations of aldehydes and ketones than n-hexanol in a counterflow diffusion flame and noted appreciable
normal alkane dopants. They also showed that the substituted amounts of 1-pentene and 1-hexene.
butanol isomers produce more propene and butene relative to n- To compare the various pentanol isomers, Tang et al. [250]
butanol, which forms more ethylene. Their flame study motivated measured ignition delay times in shock tubes at temperatures
speciation studies in non-premixed counterflow diffusion flames of above 1100 K, pressures of 1.0 and 2.6 bar, and varying equivalence
n-butanol [211,213], iso-butanol, and tert-butanol [227]. Detailed ratios. Their results showed that the ignition delay times decreased
flame structures of low-pressure premixed burner-stabilized in the order of iso-pentanol, 2-methyl-1-butanol, and n-pentanol,
flames of all four butanol isomers were measured by Obwald indicating that methyl substitution decreases reactivity. Heufer
et al. [221] using MBMS. Subsequent low-pressure premixed flame et al. [103] measured ignition delay times for n-pentanol and n-
studies with MBMS were conducted for n-butanol [218,225] and hexanol in a shock tube and RCM, spanning 9e30 bar and low to
iso-butanol [232]. These flame speciation studies provide high temperatures, and compared them to previous measurements
60 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

of alkanes. Their results showed that alcohol fuels ignite signifi- the molecular level transformation of reactants (i.e., fuel and air)
cantly faster compared to alkanes in the high temperature regime, into products via a series of elementary steps. The chemical kinetic
whereas at low intermediate temperatures, the C5 and C6 alcohols model is coupled with a physical model describing the geometry
exhibit negative temperature coefficient (NTC) behavior similar to and operating regime of the specific combustion application. A
C5 and C6 alkanes but with lower reactivity. In a recent premixed large number of differential conservation equations describing the
laminar flame speed study on 2-methyl-1-butanol, Li et al. [251] mass, momentum, energy, and species concentration are numeri-
acquired data across a remarkable range of equivalence ratios cally integrated to generate species concentration and heat release
(f ¼ 0.6e1.8), initial temperatures (393, 433, and 473 K), and initial profiles [281]. The governing equations can be solved using a nu-
pressures (1e7.5 bar), as shown in Fig. 10. Such a complete laminar merical solver that evaluates the chemical kinetic, thermodynamic,
flame speed data set at the elevated temperatures and pressures and transport properties of each differential element as time pro-
relevant to practical engines demonstrates the current state-of-the- ceeds. A commonly used program for simulating chemically
art in measuring this fundamental combustion property. reacting flow systems is Reaction Design’s CHEMKIN suite [282].
Other commonly used programs for simulating combustion include
4. Chemical kinetics of alcohol combustion CANTERA [283,284], OPENSMOKE [285,286], CMCL Innovations’
kinetics [287], and DETCHEM [288]. The CHEMKIN suite provides
Many combustion processes, including reactions in the flame modeling of a wide range of combustion instruments, including
zone resulting in heat release, reactions controlling ignition, and air shock tubes, premixed flames, diffusion flames, and partially and
pollutant formation mechanisms, occur at times when temperature perfectly stirred reactors. Chemical kinetic mechanisms are
and pressure are changing rapidly. These processes depend on the coupled with thermochemical data on all the species in the
rate of each individual chemical reaction (i.e., the reaction kinetics), mechanism to calculate forward and reverse reaction rates based
which are governed by the temperature, pressure, and the con- on the principle of microscopic reversibility. Transport properties of
centration of reactants and products. A recent review by Simmie the species are also included when attempting to model a com-
[280] describes the development and application of chemical ki- bustion process in which mixing processes, such as diffusion, are
netics models of hydrocarbon fuel combustion. A chemical kinetic rate-controlling (e.g., premixed and non-premixed flames).
model consists of a reaction mechanism, species thermochemical The process for developing and validating chemical kinetic
data, and species transport data. The reaction mechanism describes models was outlined by Frenklach et al. [289] and summarized by
Simmie [280]. Fig. 11 presents a flowchart of the model develop-
ment process. Chemical kinetic models are developed and tested
against experimental data from one or more well-characterized
combustion experiments, such as those described above. The
flowchart indicates that developing and validating a model is an
iterative process involving experiments and simulations, which are
combined to achieve a refined model. Computational chemistry
and reaction rate theory are important for the accurate determi-
nation of reaction rate, thermochemical, and transport data, while
sensitivity and reaction flux analysis help identify areas of the ki-
netic model that warrant further refinement.
The remainder of this section discusses the development of
chemical kinetic models for alcohols, specifically the formulation of
reaction mechanisms and selection of rate coefficients. Not dis-
cussed are methods for determining thermochemical properties
(e.g., group additivity [290e292] or ab initio calculations [293]), and
methods for determining transport properties (e.g., empirical cor-
relations) as described elsewhere [211,248,252,294e297]. Finally a
consistent chemical kinetic model for common C1eC5 alcohols (i.e.,
methanol, ethanol, both propanol isomers, all four butanol isomers,
and two pentanol isomers) is presented with the aim to improve
understanding of the combustion chemistry of alcohols.

4.1. The molecular structure of alcohols

Before considering comprehensive chemical kinetic models, we


should first understand the structural features of alcohols that
distinguish them from other fuels and that, in turn, affect their
molecular level transformations during combustion. Alcohols
contain a hydroxyl moiety connected to a hydrocarbon chain,
which results in unique thermochemical and reaction kinetics
properties in comparison to hydrocarbons. These effects include
phenomena such as weaker CeH bond strengths where the carbon
is bound to the hydroxyl moiety and hydrogen bonding interactions
when reacting with OH and HO2 radicals. These unique features
provide interesting discussion points in regard to the combustion
Fig. 10. Laminar flame speeds versus equivalence ratios at different temperatures and
chemistry of alcohols.
pressures for 2-methyl-1-butanol/air mixtures. The molecular structures of alcohols are routinely determined
Reprinted from Li et al. [251] with permission from Elsevier. using quantum chemistry software packages (e.g., Gaussian 09
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 61

Fig. 11. Flowchart for the development and validation of chemical kinetic models.

[298]) to identify the most stable conformers and then to coefficients for hydrogen abstraction on a per H-atom basis in-
calculate bond dissociation energies (BDEs). The CBS-QB3, G3, crease in the order of primary a < secondary a < tertiary a. The
and G4 composite methods are usually selected for their reported b-sites in the alcohols exhibit stronger than expected CeH BDEs,
accuracies of 1.0 and 1.1 kcal mol1, respectively [248,252,299e which is a consequence of the weakened adjacent a BDEs. The b
301]. Fig. 12 presents the results of such a calculation for n- CeH bonds are again ordered in decreasing strength for primary
pentanol [248] and iso-pentanol (3-methyl-1-butanol) [252] (w103 kcal mol1) < secondary (w100 kcal mol1) < tertiary
showing the lowest energy conformer and BDEs calculated us- (w98 kcal mol1) sites. The effects of the hydroxyl moiety are
ing various levels of theory. Extensive discussions on these cal- diminished beyond the b site [302], wherein the subsequent g, d,
culations and comparisons amongst various levels of theory (e.g., and ε CeH BDEs are identical to those of alkanes (e.g., primary
G4 and CBS-QB3) are presented in Refs. [248,252]. Fig. 13 presents (w102 kcal mol1) < secondary (w99 kcal mol1) < tertiary
CeH and OeH BDEs calculated at the CBS-QB3 level of theory for (w96 kcal mol1)). In the absence of any secondary effects (e.g.,
methanol, ethanol, two propanol isomers, four butanol isomers, hydrogen bonding with OH radicals), hydrogen abstraction rates
n-pentanol, and iso-pentanol [248,252]. Even considering the from the g, d, and ε sites can be estimated from established rate
1 kcal mol1 uncertainty, there are some clear trends among the rules for alkane fuels [260].
various alcohols. The OeH BDEs are strong in all alcohols, making
it difficult to abstract the hydrogen atoms. The presence of the 4.2. Formulation of comprehensive chemical kinetic models
hydroxyl moiety weakens the CeH bond at the adjacent a-site,
such that these hydrogen atoms are the easiest to abstract from The comprehensiveness of a chemical kinetic model is
all alcohols. The a-site in methanol is a primary CeH bond with measured by its ability to describe combustion phenomena
the highest BDE (w96 kcal mol1) of all the a-C-H bonds in the extensively. A model is not considered comprehensive if it has only
alcohols. The a-sites in ethanol, n-propanol, n-butanol, iso- been tested against a single experiment because the role of each
butanol, n-pentanol, and iso-pentanol are secondary CeH bonds elementary reaction varies with temperature, pressure, and
(BDE w95 kcal mol1), whereas iso-propanol and 2-butanol composition. For example, reactions between hydrogen atoms and
contain tertiary a-sites with BDE w94 kcal mol1. The rate fuel molecules are dominant in fuel-rich conditions, while

Fig. 12. Schematic diagrams of the bond dissociation energies for n-pentanol (left) and iso-pentanol (right). The numbers in the brackets are G4 values, those in parenthesis are G3
values, and those without the brackets are the CBS-QB3 values. Blue values are CeH and black are CeC BDEs (kcal mol1 at 298 K). (For interpretation of the references to color in
this figure legend, the reader is referred to the web version of this article.)
Reprinted from [248,252] with permission from Elsevier and American Chemical Society.
62 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

not available in the literature; however, such a model is needed to


highlight differences in ignition delay times, premixed laminar
flame speeds, and important combustion products. We therefore
present a consistent C1eC5 alcohol model herein, but the reader
should not consider this a “new” modeling study. Instead, this is an
assembly of data on reaction mechanisms, rate coefficients, and
species thermodynamic and transport properties available in the
literature. The complete modeling files in CHEMKIN-PRO [282]
format are available as Supplementary Material to this publica-
tion and from the KAUST website cpc.kaust.edu.sa. The chemical
kinetic model was compiled and tested using a commercial version
of CHEMKIN-PRO [282], so researchers using older non-commercial
CHEMKIN versions or other simulation tools are encouraged to
contact the corresponding author when compatibility issues hinder
the models usage.
Compiling a comprehensive model for alcohol fuel combustion
first requires the election of a certain modeling methodology. At
present, there are various methods for building a comprehensive
chemical kinetic model, but most revolve around the same prin-
ciples. Essentially, a model should include a consistent set of re-
action classes and rate coefficient estimations or “rate rules”.
Battin-Leclerc et al. [315] recently reviewed automatic generation
methods for building chemical kinetic models, all of which have
been used to develop models for alcohol fuels (e.g., CNRS Nancy’s
EXGAS by Battin-Leclerc et al. [207,316], MIT’s RMG by Green et al.
[121,233], and Milano’s MAMOX by Ranzi et al. [317,318]). In this
study, we follow the manual generation methods employed by
Curran et al. [319,320] for large alkane combustion. Some key
points described by Basevich [304] and Westbrook and Dryer [303]
are considered, such as:

 The chemical kinetic model should reproduce quantitative


Fig. 13. CeH bond dissociation energies for fuel alcohols (kcal mol1) calculated at combustion properties, such as the rate of energy release, igni-
CBS-QB3 level of theory. tion delay time, premixed laminar flame speed, and product
compositions.
reactions between hydroxyl radicals and fuel molecules dominate  The model should be detailed enough to be useful for kinetics
in fuel-lean conditions. Many reactions are important only at low investigations and explanations of combustion phenomena, for
temperatures, while others are dominant at high temperatures. In which simple schemes are not suitable.
early work on chemical kinetic mechanisms for hydrocarbon  The model should be developed in a hierarchical manner
combustion, Westbrook and Dryer [303] and Basevich [304] because combustion sequentially fragments the fuel into
explained that a comprehensive model must be compared against smaller intermediate species and eventually final products. For
experimental data covering chemically reacting flows at various example, the early work by Warnatz [321] and Dove and War-
temperatures, pressures, and reactant compositions in shock tubes natz [322] demonstrated that laminar flame speed simulations
and rapid compression machines, flow and stirred reactors, and of many hydrocarbons and methanol are primarily sensitive to
laminar flames. This approach presents a challenge for modelers the small molecule reactions H þ O2 % O þ OH,
because experimental data spanning such a wide variety of con- CO þ OH% CO2þH, and H þ O2(þM)% HO2(þM).
ditions are lacking for the higher molecular weight alcohols.  A scarcity in fundamental kinetics data presents a concern
The development of a comprehensive model for alcohol fuels because rate coefficients for many important reaction rates for
requires a basic understanding of hydrocarbon oxidation in com- alcohols are unknown. To resolve these concerns, one should
bustion systems. This was first described by Semenov [259] in his compile experimental data on measured rate coefficients, utilize
description of free radical chain reactions in combustion systems, available quantum chemistry and theoretical rate calculations,
and led to the later descriptions by Lewis and von Elbe [305e307]. and rely on estimates when necessary.
Westbrook and Dryer [303] and Basevich [304] applied these  While this approach does present uncertainties, the objective is
methods to small hydrocarbons, Miller and Bowman [308] to ni- to limit the tuning of rate coefficients to force agreement be-
trogen combustion chemistry, and Battin-Leclerc [98] to low- tween experiment and model, and to preserve the generality of
temperature alkane oxidation. Comprehensive models for various the reaction mechanism. This approach is favorable because the
alcohols exist in the literature, including Westbrook and Dryer’s mechanism can be easily updated as new reaction pathways and
early methanol model [309], Marinov’s [310] and Egolfopoulos more accurate rate coefficients (measured or calculated)
et al.’s [159] ethanol models, and Sarathy et al.’s [260] butanol iso- become available.
mer model, to name a few. Following a recent review by Tran et al.
[311], Table 4 presents a summary of comprehensive chemical ki-
netic models currently available for alcohols along with their 4.3. Reaction mechanism, reaction classes, and rate rules
experimental targets.
A single comprehensive model capable of elucidating similar- The detailed reaction mechanism for C1eC5 alcohols includes
ities and differences in the combustion of various C1eC5 alcohols is both low-temperature and high-temperature kinetic schemes for
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 63

Table 4
Comprehensive chemical kinetic models for combustion of alcohol fuels.

References Experimental targets

Methanol
Westbrook and Dryer, 1979 [309] Shock tubes, flow reactors, and laminar flame speeds
Egolfopoulos et al., 1992 [137] Propagation speed and structure of premixed laminar flames, species
concentration evolutions in flow reactors, static reactors, and shock tubes
Held and Dryer, 1998 [312] Static reactors, flow reactors, shock tubes, and premixed laminar flames
Li et al., 2007 [313] Static reactors, flow reactors, shock tubes, and premixed laminar flames

Ethanol
Egolfopoulos et al., 1992 [159] Premixed laminar flames, flow reactors, and shock tubes
Marinov, 1999 [310] Shock tubes, flow reactors, premixed laminar flames, counterflow
diffusion flames, and jet-stirred reactors
Saxena and Williams, 2007 [168] Shock tubes, premixed laminar flames, and counterflow diffusion flames
Leplat et al., 2011 [177] Shock tubes, flow reactors, premixed laminar flames, counterflow
diffusion flames, and jet-stirred reactors
Lee et al., 2012 [181] Shock tubes, rapid compression machines, and premixed laminar flames
Metcalfe et al., 2013 [314] Shock tubes, flow reactors, premixed laminar flames, counterflow
diffusion flames, and jet-stirred reactors

Propanol isomers
Frassoldati et al., 2010 [195] (n- and iso-propanol) Shock tubes, flow reactors, premixed laminar flames, and counterflow
diffusion flames

Butanol isomers
Dagaut et al., 2009 [208,211] (n-butanol) Premixed laminar flames, counterflow diffusion flames, and jet-stirred reactors
Black et al., 2010 [212] (n-butanol) Shock tubes and jet-stirred reactors
Grana et al., 2010 [213] (four isomers of butanol) Shock tubes, flow reactors, and counterflow diffusion flames
Harper et al., 2011 [121] (n-butanol) Shock tubes, flow reactors, coflow and counterflow diffusion flames, and
jet-stirred reactors
Sarathy et al., 2012 [260] (four isomers of butanol) Shock tubes, rapid compression machines, premixed laminar flames, and
jet-stirred reactors
Merchant et al., 2013 [233] (iso-butanol) Shock tubes, flow reactors, premixed laminar flames, coflow and counterflow
diffusion flames, rapid compression machines, and jet-stirred reactors

Higher alcohols
Togbé et al., 2010 (n-hexanol) Premixed laminar flames and jet-stirred reactors
Togbé et al., 2011 [246] (n-pentanol) Premixed laminar flames and jet-stirred reactors
Heufer et al., 2012 [248] (n-pentanol) Shock tubes, rapid compression machines, premixed laminar flames, and jet-stirred reactors
Sarathy et al., 2013 [252] (iso-pentanol) Shock tubes, rapid compression machines, jet-stirred reactors, premixed laminar
flames, and non-premixed counterflow diffusion flames

alcohol fuels. The mechanism is compiled in a hierarchical alcohol combustion are omitted in the present manuscript. Readers
manner, starting with the oxidation of small hydrocarbons and are referred to the Supplemental Material of [260] for a more
alcohols, and then sequentially adding reaction mechanisms for detailed discussion. Fig. 14 presents a generalized scheme for the
larger alcohols. The base C0eC2 hydrocarbon oxidation mechanism primary mechanism of alcohol oxidation, to which the pyrolysis
used here is the recently published AramcoMech 1.3 developed at reactions are added for a complete mechanism. Methanol and
the National University of Ireland Galway (NUIG) by Metcalfe, ethanol do not exhibit low-temperature reactivity under typical
Burke, and Curran with Ahmed from Saudi Aramco [314]. Their H2/ combustion conditions, so including comprehensive low-
CO/O2 sub-mechanism is based on the work of Kéromnès et al. temperature reaction pathways is not required. Table 5 presents
[323]. The C4 and C5 alkane and alkene sub-mechanisms were experimentally measured reaction rate coefficients available in the
taken from Healy et al. [324]. The methanol and ethanol sub- literature for various alcohol-related reactions. This table is cited
mechanisms are from the recent work of Metcalfe et al. [314] along with references to theoretical studies to highlight important
and Mittal et al. [187], respectively. The n- and iso-propanol sub- rate coefficient studies, existing discrepancies, and areas for future
mechanism is from Johnson et al. [193] with several modifica- work. The reaction classes included for alcohol oxidation are shown
tions explained by Man et al. [200] plus several important low in Box 1, approximately in sequence along the oxidation reaction
temperature reaction pathways. For the higher alcohols, including from fuel to fully oxidized products.
the four butanol isomers and two pentanol isomers (n-pentanol
and 3-methyl-1-butanol), few models [233,248,249,252,260] are 4.3.1. Reaction class 1: unimolecular alcohol decomposition
comprehensive, such that they can be used to simulate both low- At high temperatures (above 1500 K) and under fuel-rich or
temperature and NTC reactivity (e.g., ignition) and high- pyrolysis conditions, the consumption of alcohols is dominated by
temperature combustion phenomena (e.g., laminar flame unimolecular decomposition. Such reactions are important for
speeds). Thus, the mechanisms for butanol isomers by Sarathy correctly simulating soot emissions in fuel-rich combustion, igni-
et al. [260], n-pentanol by Heufer et al. [248], and iso-pentanol (3- tion delay times in highly diluted environments, and non-premixed
methyl-1-butanol) by Sarathy et al. [252] are chosen for the pre- diffusion flame structures. In all alcohols, these reactions consist of
sent analysis. simple CeC and CeH bond scission reactions. Unimolecular water
The important reaction classes are discussed in greater detail elimination (i.e., dehydration) reactions are also important for al-
below. As noted previously, these reaction classes follow from cohols. Moc et al. [325] explained that these reactions proceed via
methodologies initially established by Westbrook and Dryer stretching of the CeO bond with a concomitant stretching and
[303,309] and later applied to n-heptane modeling by Curran et al. breaking of a CeH bond (a, b, g, d, etc.). They proceed via three-
[319]. Several reaction classes that are not unique or important for center (a carbon), four-center (b), five-center (g), or larger
64 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

are of relevance to atmospheric chemistry, or at high temperatures


above 1000 K. H-atom abstraction reactions by HO2 radicals have
not been directly measured for any of the alcohols. The lack of
direct measurements across the range of combustion temperatures
thus requires application of computational theory and estimation
methods to determine the appropriate rate coefficients for reaction
mechanisms [326].

4.3.3. Reaction classes 3 and 4: alcohol radical decomposition and


alcohol radical isomerization
Fuel radicals can undergo unimolecular decomposition via b-
scission at high temperatures (e.g., above 900 K). In these reactions,
the bond once removed from the radical site (i.e., b to the radical
site) break to form an unsaturated species and another radical. b-
scission of CeC and CeH bonds leads to the formation of unsatu-
rated alcohols or alkenes (i.e., C]C bonds), whereas b-scission of
the OeH bonds results in the formation of an aldehyde (i.e., C]O
bond). Fuel radicals can also isomerize by transferring H-atoms
from any carbon/hydroxyl site to the radical site. The rate coeffi-
cient for these reactions depends on the nature of the CeH bonds
broken and formed (i.e., primary, secondary, tertiary, or hydroxyl)
and on the ring strain energy barrier. Typically, radical isomeriza-
tions involving 5-, 6-, or 7-member transition state rings are
considered due to their lower ring strain energies.
Fig. 14. Simplified scheme for the primary mechanism of alcohol oxidation of alcohols.
Figure adapted from Battin-Leclerc [98].
4.3.4. Reaction classes 5e9: reactions involving unsaturated
intermediates (enols and carbonyls)
transition state rings. The four-center reactions involving the b Carbonyls are formed in appreciable quantities during alcohol
hydrogen atoms are the most thermodynamically favored and combustion. At high temperatures, H-atom abstraction and b-
result in the formation of an alkene and water, as shown in Fig. 15. scission of parent alcohols can result in the formation of alde-
The rates of these reactions can be determined experimentally, and hydes/ketones. At lower temperatures, the reaction of a-hydrox-
several important studies are presented in Table 5. The rate co- yalkyl radicals with O2 leads to the direction formation of
efficients of four-center dehydration reactions vary with the extent carbonyls. Typical carbonyl intermediates include formaldehyde
of substitution because this affects the number of abstractable b (methanal), acetaldehyde (ethanal), propionaldehyde (n-propa-
hydrogen atoms and their BDEs (i.e., primary, secondary, or ter- nal), acetone, butyraldehyde (butanal), and all possible isomers for
tiary). The structure of the product alkene affects the global reac- C3 and larger carbonyls. A thorough review of aldehyde/ketone
tivity of the alcohol. combustion is not available in the literature, and preparing one is
beyond the scope of this manuscript. Highlighted briefly are
several recent studies on carbonyl combustion experiments and
4.3.2. Reaction class 2: H-atom abstraction from the alcohol modeling. The destruction pathways of formaldehyde have been
H-atom abstraction reactions are the primary means by which extensively studied because all hydrocarbons are converted to CO
fuel alcohols are consumed in combustion environments. A variety through formaldehyde. Li et al. [313] presented a comprehensive
of small radical species can abstract hydrogen atoms from alcohols chemical kinetic modeling study of formaldehyde combustion, and
(e.g., H, OH, HO2, O, CH3O2, CH3O, CH3, C2H5, etc.). The most also reviewed much of the previous experimental and theoretical
important radicals are OH and HO2 radicals at low and intermediate studies. Recent studies on combustion of propanal [327] and
temperatures and H radicals at high temperatures and/or under butanal isomers [328] were conducted by Veloo et al., and they
fuel-rich conditions; only rate coefficients and branching ratios for present a complete discussion of previous works on these fuels. In
these will be discussed in this review. In alcohols, the H atoms can general, aldehydes are consumed via hydrogen atoms abstracted
be characterized as hydroxyl, a, b, g, etc., as well as primary, sec- from the weakly bound aldehydic HeC]O site, following by a-
ondary, or tertiary (Fig. 13). These reactions are typically written in scission to form CO and an alkyl radical. Larger aldehydes can
the forward direction (i.e., H/OH/HO2 þ fuel), and the reverse rate exhibit low-temperature reactivity via oxidation of their alkyl in-
coefficients are calculated based on the principle of microscopic termediates [327,328], as shown in Fig. 16.
reversibility. The rate coefficient for abstraction depends on the Enols are common intermediates during the combustion of al-
radical species and the type of H atom being abstracted. As shown cohols via dehydrogenation. The isomerization of enols to alde-
previously, hydrogen bound in the OH moiety is the most difficult hydes or ketones (i.e., enoleketo tautomerization) are slow in the
to abstract, followed by primary H atoms, while tertiary H atoms gas-phase due to large energy barriers (w56 kcal/mol) [329];
are the weakest and most easily abstracted. a CeH bonds are however, these reactions occur rapidly in the condensed phase or
weaker due to the nearby presence of the OH group, whereas b CeH on surfaces [212]. The ketone/aldehyde tautomer is typically more
bonds are slightly stronger than analogous bonds in an alkane stable than the enol form, such that reaction mechanisms need to
molecule. include all realistic pathways for such tautomerization reactions.
Accurate temperature-dependent rate coefficients and branch- Enoleketo tautomerizations and isomerizations catalyzed by H
ing ratios for H-atom abstraction by OH radicals are especially radicals [218], as well as isomerizations catalyzed by HO2 [330] are
important for simulating alcohol reactivity from 500 to 1500 K. As important. The H radical catalyzed isomerization can proceed via
shown in Table 5 direct rate measurements of these reactions have an a-hydroxyalkyl intermediate species (e.g.,
been performed either at low temperatures (e.g., <400 K), which C2H3OH þ H % sC2H4OH; sC2H4OH % CH3CHO þ H). Additionally,
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 65

Table 5
Overview of recent experimental studies of elementary reaction kinetics under conditions relevant to ignition and combustion of alcohols. Low-temperature atmospheric
studies below 400 K are included if they provide kinetic and/or mechanistic information. This table does not include experimental studies related to CH3OH and C2H5OH prior
to 2001, which have already been reviewed and presented by Baulch et al. [352]

References Reaction system Method of initiation Detection technique T, P range

Methanol
H-atom abstraction
Jimenez et al., 2003 [353] CH3OH þ OH Laser photolysis LIF 235e360 K
Baulch et al., 2005 [352] CH3OH þ OH Reviewed literature
prior to 2001
Dillon et al., 2005 [354] CH3OH þ OH Laser photolysis LIF 210e351 K
Srinivasan et al., 2007 [355] CH3OH þ OH Shock tube thermal EAD 1591e1710 K
Baulch et al., 2005 [352] CH3OH þ H Reviewed literature
prior to 2001. No recent
experimental work.
Lu et al., 2005 [356] CH3OH þ O Shock tube thermal ARAS 835e1777 K
Baulch et al., 2005 [352] CH3OH þ O Reviewed literature
prior to 2001
Carr et al., 2011 [357] CH3OH þ O Laser photolysis LIF 775 K
Baulch et al., 2005 [352] CH3OH þ CH3 Reviewed literature
prior to 2001
Park et al., 2012 [358] CH3OH þ phenyl Laser photolysis CRDS and MS 85e1777 K

Unimolecular decomposition
Baulch et al., 2005 [352] CH3OH to products Reviewed literature
prior to 2001
Srinivasan et al., 2007 [355] CH3OH to products Shock tube thermal EAD 1591e2865 K
Lu et al., 2010 [359] CH3OH to products Shock tube thermal ARAS 13591644 K
Lee et al., 2013 [360] CH3OH to products Shock tube thermal ARAS 304e767 K

Radical decomposition
Baulch et al., 2005 [352] CH2OH to products Reviewed literature
prior to 2001

Ethanol
H-atom abstraction
Meier et al., 1985 [361] C2H5OH þ OH Discharge flow reactor LIF 300e1000 K
Greenhill and Ogrady, C2H5OH þ OH Laser photolysis RA 255e459 K, 1.0133 bar
1986 [362]
Wallington and Kurylo, C2H5OH þ OH Laser photolysis RF 240e440 K, 0.0333e0.0666 bar
1987 [363]
Hess and Tully, 1988 [364] C2H5OH þ OH Laser photolysis LIF 293e750 K, 0.9333 bar
Bott and Cohen, 1991 [365] C2H5OH þ OH Shock tube thermal UV laser absorption w1200 K, 1.01325 bar
Jimenez et al., 2003 [353] C2H5OH þ OH Laser photolysis LIF 227e360 K, 0.0533 bar
Dillon et al., 2005 [354] C2H5OH þ OH Laser photolysis LIF 216e368 K, 0.1332 bar
Sivaramakrishnan C2H5OH þ OH Shock tube thermal EAD 857e1297 K
et al., 2010 [366]
Carr et al., 2011 [367] C2H5OH þ OH Laser photolysis LIF 298e900 K, 0.0067e0.1333 bar
Orkin et al., 2011 [368] C2H5OH þ OH Laser photolysis RF 220e370 K, 0.0133e0.2666 bar
Stranic et al., 2014 [369] C2H5OH þ OH Shock tube thermal UV laser absorption 900e1270 K, w1 bar
Sivaramakrishnan C2H5OH þ H Shock tube thermal ARAS 1054e1359 K
et al., 2010 [366]
Carr et al., 2011 [357] C2H5OH þ O Laser photolysis LIF 650e860 K
Park et al., 2012 [358] C2H5OH þ phenyl Laser photolysis CRDS and MS 835e1777 K

Unimolecular decomposition
Li et al., 2004 [370] C2H5OH to products Flow reactor FTIR 1045e1080 K, 1.7225e3.0398 bar
Baulch et al., 2005 [352] C2H5OH to products Reviewed literature
prior to 2001
Sivaramakrishnan C2H5OH to products Shock tube thermal ARAS 1308e1732 K, 0.20e0.709 bar
et al., 2010 [366]
Wu et al., 2011 [371] C2H5OH to products Shock tube thermal ARAS 1450e1760 K, 1e2 bar

Propanol isomers
Wallington and iso-C3H7OH þ OH Laser photolysis RF 240e440 K, 0.0333e0.0667 bar
Kurylo, 1987 [363]
Dunlop and Tully, 1993 [372] iso-C3H7OH þ OH Laser photolysis LIF 293e745 K, w1 bar
Yujing and Mellouki, n-C3H7OH þ OH Laser photolysis LIF 253e372 K, 0.0399e0.3999 bar
2001 [373]
Yujing and Mellouki, iso-C3H7OH þ OH Laser photolysis LIF 253e372 K, 0.0399e0.3999 bar
2001 [373]
Rajakumar et al., 2010 [374] n-C3H7OH þ OH Laser photolysis LIF 327e376 K, 0.1067e0.5333 bar
Rajakumar et al., 2010 [374] iso-C3H7OH þ OH Laser photolysis LIF 327e376 K, 0.1067e0.5333 bar
Orkin et al., 2012 [375] iso-C3H7OH þ OH Laser photolysis RF 240e370 K, 0.0399e0.1333 bar

Butanol isomers
H-atom abstraction
Wallington et al., 1988 [376] tert-C4H9OH þ OH Laser photolysis Fluorescence 240e440 K, 0.0333 and 0.0667 bar
Teton et al., 1996 [377] tert-C4H9OH þ OH Laser photolysis LIF 253e372 K, 0.1333 bar
(continued on next page)
66 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Table 5 (continued )

References Reaction system Method of initiation Detection technique T, P range

Yujing and Mellouki, n-C4H9OH þ OH Laser photolysis LIF 253e372 K, 0.0399e0.3999 bar
2001 [373]
Mellouki et al., 2004 [378] iso-C4H9OH þ OH Laser photolysis LIF 241e373 K, 0.1333 bar
Jimenez et al., 2005 [379] 2-C4H9OH þ OH Laser photolysis LIF 263e354 K, 0.0547e0.2573 bar
Vasu et al., 2010 [380] n-C4H9OH þ OH Shock tube thermal UV laser absorption 1017e1269 K, 2.279 bar
Pang et al., 2012 [381] n-C4H9OH þ OH Shock tube thermal UV laser absorption 900e1200 K, w1 bar
Pang et al., 2012 [382] iso-C4H9OH þ OH Shock tube thermal UV laser absorption 907e1147 K, w1 bar
Pang et al., 2012 [383] 2-C4H9OH þ OH Shock tube thermal UV laser absorption 888e1178 K, w1 bar
Stranic et al., 2013 [384] tert-C4H9OH þ OH Shock tube thermal UV laser absorption 900e1200 K, w1.1 bar
McGillen et al., 2013 [385] n-C4H9OH þ OH Laser photolysis LIF 221e381 K, 0.067e0.267 bar
McGillen et al., 2013 [385] 2-C4H9OH þ OH Laser photolysis LIF 221e381 K, 0.067e0.267 bar
McGillen et al., 2013 [385] iso-C4H9OH þ OH Laser photolysis LIF 221e381 K, 0.067e0.267 bar
McGillen et al., 2013 [385] tert-C4H9OH þ OH Laser photolysis LIF 221e381 K, 0.067e0.267 bar

Isomerization
Heiss and Sahetchian, 1-C4H9O Flow reactor thermal GC/MS 34e503 K, w1 bar
1996 [349]

Unimolecular decomposition
Tsang, 1964 [203] tert-C4H9OH to products Shock tube thermal GC (detector not specified) 10501300 K, 0.1067e0.4667 bar
Lewis et al., 1974 [204] tert-C4H9OH to products Shock tube thermal GC (detector not specified) 9201175 K, 0.4933e2.0798 bar
Kalra et al., 2004 [386] tert-C4H9OH to products Shock tube thermal GC (detector not specified) 7091156 K, 0.07e1.6 bar
and static reactor and NMR
Rosado-Reyes and n-C4H9OH to products Shock tube thermal GC/FID/MS 11261231 K, 1.3e6.5 bar
Tsang, 2012 [387]
Rosado-Reyes and 2-C4H9OH to products Shock tube thermal GC/FID/MS 10451221 K, 1.5e6 bar
Tsang, 2012 [388]
Rosado-Reyes and iso-C4H9OH to products Shock tube thermal GC/FID/MS 1090e1240 K, 1.5e6 bar
Tsang, 2013 [389,390]

Higher alcohols
H-atom abstraction
Wallington et al., 1988 [391] 3-Methyl-2-butanol þ OH Laser photolysis RF 298 K, 0.0333e0.0667 bar
Wallington et al., 1988 [391] 2-Pentanol þ OH Laser photolysis RF 298 K, 0.0333e0.0667 bar
Wallington et al., 1988 [391] 3-Pentanol þ OH Laser photolysis RF 298 K, 0.0333e0.0667 bar
Wallington et al., 1988 [391] n-Hexanol þ OH Laser photolysis RF 298 K, 0.0333e0.0667 bar
Wallington et al., 1988 [391] 2-Hexanol þ OH Laser photolysis RF 298 K, 0.0333e0.0667 bar
Wallington et al., 1988 [391] n-Heptanol þ OH Laser photolysis RF 298 K, 0.0333e0.0667 bar
Mellouki et al., 2004 [378] 3-Methyl-1-butanol þ OH Laser photolysis LIF 241e373 K, 0.1332 bar
Mellouki et al., 2004 [378] 3-Methyl-2-butanol þ OH Laser photolysis LIF 241e373 K, 0.1332 bar
Jimenez et al., 2005 [379] 2-Methyl-2-butanol þ OH Laser photolysis LIF 263e354 K, 0.0547e0.2573 bar
Jimenez et al., 2005 [379] 2,3-Dimethyl-2-butanol þ OH Laser photolysis LIF 263e354 K, 0.0547e0.2573 bar

these reactions can be chemically activated, such that pressure [336,337] and iso-butanol [338]. Despite the known importance of
dependent rate coefficients are necessary [218,331,332]. this reaction pathway, there is very little information regarding
Reactions of an enol þ HO2 forming an aldehyde þ HO2 are the rate coefficient at combustion relevant temperatures. Recently,
based on the acetaldehyde/ethenol work of da Silva and Bozzelli Zádor et al. [339] and da Silva et al. [340] conducted studies on the
[330]. These reactions are faster pathways of enol-keto isomeriza- a-hydroxyethyl þ O2 reaction system. They showed that the
tion given suitably high levels of HO2. These reactions are impor- aforementioned reaction proceeds through an activated a-hy-
tant for enols (i.e., C]CeOH). Enol-keto isomerization catalyzed by droxy-ethylperoxy adduct that rapidly decomposes to
carboxylic acids (e.g., formic acid), which is an interesting reaction acetaldehyde þ HO2. Furthermore, the decomposition reaction
mechanism recently proposed by da Silva [333], may also be occurs with such a low barrier that conventional low-temperature
important in combustion systems. By means of theoretical calcu- chain branching pathways are unimportant, as shown in Fig. 18.
lations, he explains that formic acid reacts with vinyl alcohol (i.e., This reaction pathway essentially inhibits the low-temperature
ethenol) to form acetaldehyde via a double hydrogen shift mech- ignition behavior of alcohols. This reaction class should be
anism, as shown in Fig. 17. The proposed reaction is shown to be fast included at all temperatures because the reaction of a-hydrox-
at low temperatures and with low concentrations of formic acid yalkyl with O2 can be competitive with b-scission reactions up to
and is therefore important under atmospheric conditions [333]. 1000 K (e.g., premixed flames, intermediate-temperature shock
Other important reactions for enols include H-atom abstraction tube, etc.).
reactions, addition of H radicals to double bonds, unimolecular
decomposition of enols, and enol radical decomposition reactions 4.3.6. Reaction class 11: addition of O2 to alcohol radicals
[334,335] (R þ O2 % ROO)
The addition of molecular oxygen to fuel radicals can form sta-
4.3.5. Reaction class 10: reaction of O2 to a-hydroxyalkyl radicals to bilized hydroxyalkyl peroxy radicals (ROO), and thus initiate the
directly form an aldehyde/ketone þ HO2 low-temperature chain branching process. These reactions are
The reaction pathway involving the reaction of molecular ox- important in C4 and larger alcohols, but their rates have not been
ygen to a-hydroxyalkyl radicals to directly form an aldehyde/ke- investigated in detail. From analogy with alkanes [319,320], the rate
tone and HO2 is known to occur rapidly in the atmosphere. Some of O2 addition to a fuel radical may depend on whether the radical
notable studies of this reaction include those in n-butanol carbon is a primary, secondary, or tertiary site. Additional work is
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 67

Box 1
Reaction classes for alcohol oxidation

High-temperature reaction classes

1. Unimolecular alcohol decomposition


2. H-atom abstraction from the alcohol
3. Alcohol radical decomposition
4. Alcohol radical isomerization
5. H-atom abstraction reactions from unsaturated alcohols
(e.g., enols)
6. Enol-Keto tautomerizations and isomerizations
7. Addition of H radicals to unsaturated alcohols
8. Unsaturated alcohol radical decomposition
9. Unimolecular decomposition of unsaturated alcohols
10. Reaction of O2 to a-hydroxyalkyl radicals to directly
form an aldehyde/ketone þ HO2

Low-temperature reaction classes (R refers to a alcohol


_
radical such as CH3CHCH 2CH2OH)

11. Addition of O2 to alcohol radicals (R þ O2 % ROO)


12. R þ ROO % RO þ RO
13. R þ HO2 % RO þ OH
14. R þ CH3O2 % RO þ CH3O
15. ROO radical isomerization (ROO % QOOH) including
Waddington type reaction mechanism
16. Concerted eliminations (ROO % unsaturated
alcohol þ HO2)
17. ROO þ HO2 % ROOH þ O2
18. ROO þ H2O2 % ROOH þ HO2
Fig. 15. Four-center dehydration reactions for alcohol fuels.
19. ROO þ CH3O2 % RO þ CH3O þ O2
20. ROO þ ROO % RO þ RO þ O2
21. ROOH % RO þ OH
needed to determine rate coefficients for this reaction class using 22. RO decomposition
quantum chemical potential energy calculations and master 23. Formation of epoxy alcohols via cyclization
equation methods. 24. QOOH % unsaturated alcohol þ HO2 (radical site b to
OOH group)
4.3.7. Reaction classes 12e14: R þ ROO % RO þ RO, 25. QOOH % alkene/unsaturated alcohol þ carbonyl þ OH
R þ HO2 % RO þ OH, and R þ CH3O2 % RO þ CH3O (radical site g to OOH group)
Reactions classes 12, 13, and 14 are grouped together because 26. Addition of O2 to QOOH (QOOH þ O2 % OOQOOH)
they are analogous reactions. Each reaction proceeds through a 27. Reaction of O2 with a-hydroxyalkyl hydroperoxy radi-
chemically activated ROOR0 adduct which decomposes to RO þ R0 O cals (e.g., CH3CH(OOH)CH2CHOH_ þ O2)
(where R0 ¼ R, CH3, H). The reactions are analogous to 28. Isomerization of OOQOOH and formation of carbonyl
CH3 þ HO2 % CH3O þ OH [341]. hydroxyalkyl hydroperoxides and OH, including
Waddington-type reaction mechanisms
4.3.8. Reaction class 15: ROO radical isomerization (ROO % QOOH) 29. Decomposition of carbonyl hydroxyalkyl hydroperox-
including Waddington type reaction mechanism ides to form carbonyl radical species and OH
The second step in the low-temperature radical chain branch- 30. Epoxy alcohols reactions with OH and HO2
ing process is the intramolecular H-abstraction of hydroxyalkyl
peroxy radicals (ROO) to form hydroxyalkyl hydroperoxy radicals
(QOOH). The most important isomerizations include 5-member, 6-
member, and 7-member transition state rings. For alcohols, the chain branching pathways, thereby inhibiting low-temperature
reaction rate coefficients for these reactions are not well studied, reactivity. Sun et al. [344] conducted theoretical calculations on
so estimations techniques are required. The rate coefficient de- these Waddington reactions for tert-butanol and iso-butanol sys-
pends on the nature of the broken CeH bond (i.e., primary, sec- tems. Their study suggests that the b-ROO radical almost
ondary, or tertiary) and on the ring strain energy barrier [260].
Waddington type reaction pathways [342,343] are known to be a
characteristic part of alkene oxidation, but are part of alcohol
oxidation too. These reactions involve the b-ROO radical under-
going a six-membered ring isomerization to abstract an H atom
from the hydroxyl moiety (i.e., forming an alkylhydroperoxide Fig. 16. Reaction path of propanal low-temperature oxidation in a JSR, 10 atm, 655 K,
alkoxy), followed by a rapid decomposition. These reactions are stoichiometric conditions.
chain propagating and compete directly with low-temperature Reprinted from Veloo et al. [327] with permission from Elsevier.
68 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

depend on the nature of both the CeOO bond and the CeH bond
broken (i.e., primary, secondary, or tertiary) during the direct
elimination reaction; however, there is a lack of fundamental
experimental and theoretical rate parameters studies on this re-
action class for alkanes, let alone alcohols, to develop such a
rigorous rate rule.

4.3.10. Reaction classes 17e20: ROO þ HO2 % ROOH þ O2,


ROO þ H2O2 % ROOH þ HO2, ROO þ CH3O2 % RO þ CH3O þ O2,
and ROO þ ROO % RO þ RO þ O2
ROO can undergo bimolecular reactions with various radicals
(e.g., HO2, CH3O2, ROO) to form ROOH, which readily decomposes to
RO and OH radicals (i.e., class 21). ROO can also react with H2O2 to
form another stable species and a peroxy species. This reaction
Fig. 17. Transition-state structure for the double hydrogen shift reaction of vinyl sequence interconverts H2O2 to ROOH, which can decompose at
alcohol with formic acid to form acetaldehyde. relatively lower temperatures to produce OH radicals. These re-
Reprinted from da Silva et al. [333] with permission from John Wiley and Sons. actions are discussed in further detail in the previous alkane work
of Curran et al. [319,320].

exclusively reacts via the Waddington mechanism. Recently, 4.3.11. Reaction classes 21 and 22: ROOH % RO þ OH and RO
Welz et al. [345,346] described Waddington-type pathways decomposition
involving hydrogen transfer from the OH group as important for The decomposition ROOH species leads to the formation of two
the g-ROO and d-ROO radicals via 7- and 8-membered transition reactive radical species. This reaction pathway is generally unim-
state rings. portant because the pathways forming ROOH from RO2 (i.e., classes
17 and 18) are in competition with the RO2 % QOOH isomerization
4.3.9. Reaction class 16: concerted eliminations reaction pathways (i.e., class 15). However, it cannot be completely
(ROO % enol þ HO2) neglected because the concerted elimination pathway (i.e., class 16)
The concerted (direct) elimination of HO2 from the ROO radical increases the HO2 concentration, which enhances the
occurs via a 5-membered transition state and was first discovered ROO þ HO2 % ROOH þ O2 (i.e., class 17) pathway. The b-scission of
in alkanes by Quelch et al. [347]. The same type of reactions is RO radicals leads to production of an alkyl/hydroxyalkyl radical plus
expected to occur in alcohols. Zádor et al. [99] explain that the OeO an aldehyde/ketone/acid.
moiety leaves the ROO molecule, and takes an H atom from the
adjacent C atom with it. This reaction pathway competes directly 4.3.12. Reaction class 23: formation of epoxy alcohols via
with the ROO radical isomerization pathway to QOOH. Therefore cyclization
the concerted elimination path competes with the chain branching The cyclization of a QOOH radical to form an epoxy alcohol (i.e.,
path and reduces overall low-temperature reactivity. It is also analogous to a cyclic ether in alkane oxidation) plus an OH radical is
responsible for the majority of HO2 and enols formed in the low- an important pathway that competes with the critical chain
and intermediate-temperature regimes. The rate coefficient may branching channel QOOH þ O2 % OOQOOH (i.e., class 25). The
activation energies and rate coefficients for these reactions are not
well known for alcohols, so alkane data [319] can be used as an
estimate. This particular reaction is minor, as cyclic epoxy alcohols
have not yet been measured in significant quantities during alcohol
combustion [345,346].

4.3.13. Reaction classes 24 and 25: QOOH % enol þ HO2 (radical


site b to OOH group) and QOOH % alkene þ carbonyl þ OH (radical
site g to OOH group)
The decomposition of a QOOH with a radical site b to the
hydroperoxy group leads to the formation of an enol plus HO2.
The decomposition of a QOOH when the radical site is g to the
OOH group forms an alkene, an aldehyde or ketone, and an OH
radical.

4.3.14. Reaction class 26: addition of O2 to QOOH


(QOOH þ O2 % OOQOOH)
The addition of molecular oxygen to hydroxyalkyl hydroperox-
ide radicals (i.e., QOOH) is the third step in the low-temperature
chain branching process. The rate of O2 addition to a QOOH
radical depends on whether the radical carbon is an a, primary,
secondary, or tertiary site.

4.3.15. Reaction class 27: reaction of O2 with a-hydroxyalkyl


Fig. 18. Rate coefficients (k, cm3 mol1 s1) as a function of temperature and pressure _
for formation of acetaldehyde þ HO2 (black lines), vinyl alcohol þ HO2 (gray lines), and
hydroperoxy radicals (e.g., CH3CH(OOH)CH2CHOH þ O2)
stabilized C2H5O3 radicals (white lines), in the reaction of a-hydroxyethyl þ O2. The importance of molecular oxygen reaction with a a-
Reprinted from da Silva et al. [340] with permission from ACS Publications. hydroxyalkyl radical to form an aldehyde/ketone and HO2 was
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 69

previously discussed (i.e., class 10). This reaction is important in from 500 to 1500 K, pressures greater than 10 bar, and equiv-
inhibiting the low-temperature reactivity of alcohols. It has been alence ratios ranging from lean to rich (whenever available), as
shown experimentally under atmospheric conditions [348,349] these are the most relevant to internal combustion engines.
and theoretically [350,351] that an analogous reactions occurs  Simulating the concentrations of intermediate and product
for O2 reaction to a-hydroxyalkyl hydroperoxy radicals (e.g., species is important for elucidating reaction pathways of alcohol
CH3CH(OOH)CH2CH.OH þ O2), thus forming a carbonyl hydroper- fuels and simulating (potentially harmful) emissions from
oxide plus HO2. In the n-butanol system, the eventual products of alcohol-fueled engines. Thus, comparisons are presented for
this reaction path are a hydroperoxide butyraldehyde plus HO2 data obtained in premixed flames, stirred reactors, and/or flow
[348,349]. The hydroperoxide butyraldehyde decomposes to give reactors at temperatures from 500 K and above, elevated pres-
back OH; therefore, this reaction path is a form of low- sures, and a range of equivalence ratios.
temperature chain branching which produces one OH and one  Data obtained in fuel/air (21% O2 and 79% N2) mixtures are
HO2 radical. favored over data obtained with other diluents (e.g., Ar). Several
exceptions exist, such as premixed flame speciation studies
4.3.16. Reaction class 28: isomerization of OOQOOH and formation where Ar dilution is used as an internal standard for
of carbonyl hydroxyalkyl hydroperoxides and OH including calibration.
Waddington type reactions mechanism  Studies under non-premixed (e.g., counterflow flames, coflow
In this reaction class an OOQOOH radical isomerizes, releases an flames, etc.) and pyrolysis conditions are relevant to CI and
OH, and then forms a carbonyl hydroxyalkyl hydroperoxide species premixed charge compression ignition (PCCI) engine combus-
(i.e., analogous to carbonyl alkyl hydroperoxide in alkane oxida- tion and pollutant formation. Experimental studies and model
tion). This is the fourth step in the low-temperature chain simulations under such conditions are selectively presented to
branching process. The rate coefficients for these reactions are develop a mechanistic understanding of alcohol combustion in
estimated based on analogous reaction in alkane combustion fuel-rich environments.
[319,320]. An important underlying assumption in this methodol-  When available, experimental data from newer studies are
ogy is that the hydrogen atom being abstracted is bound to the favored over older data, unless the literature review has deemed
carbon atom bonded to the hydroperoxide (OOH) group, which has the newer data inaccurate.
a lower CeH bond strength than a normal CeH bond and makes it  Comparing the combustion chemistry features across the
easier to abstract. various alcohols is facilitated by utilizing data obtained from the
same experimental facilities under similar conditions.
4.3.17. Reaction class 29: decomposition of carbonyl hydroxyalkyl  All the simulations were conducted in CHEMKIN-PRO [282]
hydroperoxides to form oxygenated radical species and OH using the appropriate reactor modules. Premixed laminar
The decomposition of carbonyl hydroxyalkyl hydroperoxides flame simulations accounted for thermal diffusion (i.e., Soret
forms an OH radical, a smaller oxygenated radical, and a stable effect), assumed mixture-averaged transport, and the solutions
oxygenate (i.e., aldehyde or ketone). This reaction pathway is the were highly resolved with approximately 200 grid points (GRAD
final step in the low-temperature chain branching process because 0.1, CURV 0.1). Burner-stabilized premixed flames were simu-
it forms two radical species from one stable reactant. lated using the experimentally measured temperature profile as
a boundary condition.
4.3.18. Reaction class 30: epoxy alcohols reactions with OH and HO2
The epoxy alcohols formed via cyclization of a hydroxyalkyl
hydroperoxy radical (i.e., reaction class 23) have an oxygen atom 4.5. The combustion chemistry of methanol
embedded within the ring. One can assume these species react via
H-atom abstraction by OH and HO2, which are the predominant Methanol is the simplest alcohol and its oxidation mechanism
radicals at low and intermediate temperatures. The subsequent provides the basis for the oxidation of larger alcohols. In hydro-
epoxy alcohol radicals then undergo b-scission to form smaller carbon combustion, the recombination of methyl and hydroxyl
hydrocarbon and carbonyl species. radicals forms methanol (e.g., CH3þOH%CH3OH), and many in-
termediate species important to methane combustion (e.g., CH2O,
4.4. Model testing and analysis CH3O, CH2OH, and HCO) are also relevant to methanol combus-
tion. Notable comprehensive reaction mechanisms described for
Following the extensive discussion on reaction classes, testing of methanol are those by Westbrook and Dryer [309], Egolfopoulos
the C1eC5 chemical kinetic model against experimental data is et al. [137], Held and Dryer [312], Li et al. [313], Ing et al. [392],
necessary. Comparisons of each alcohol oxidation sub-mechanism and Aranda et al. [153]. Table 5 lists manuscripts describing the
with experimental data are presented to demonstrate its perfor- direct rate measurement of methanol decomposition and impor-
mance across a range of temperatures, pressures, and mixture tant branching fractions. Srinivasan et al. [355] utilized the re-
fractions of relevance to practical combustion devices. The goal flected shock tube technique with multipass absorption
here is not to include comparisons against all available experi- spectrometric detection of OH radicals to measure the rate co-
mental data for a specific alcohol fuel, but instead to follow efficients of important product channels between 1591 and
guidelines for selecting the targets to be presented in this review 2865 K. More recently, Lee et al. [360] and Lu et al. [359]
article, which are as follows: measured methanol thermal decomposition rates behind re-
flected shock waves by atomic resonance absorption spectrometry
 Premixed laminar flame speed simulations are important for of H atoms from 1660 to 2050 K and 1359 to 1644 K, respectively.
flame propagation in SI engines; these are presented for all the Pressure-dependent rate coefficients using high-level ab initio
fuels. Comparisons are shown for a range of equivalence ratios calculations and master equation simulations by Jasper et al. [393]
and, if data are available, at elevated pressures. are in good agreement with the aforementioned direct experi-
 Homogenous ignition delay time simulations are important for mental measurements. Jasper et al. [393] found that the branching
SI, CI, and LTC engines; these are presented for all the fuels. ratio largely favors the (R1) channel (w80%) with the (R2) channel
Comparisons are presented for data obtained temperatures (w20%) being the second most important. The important
70 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

reaction channels and pressure-dependent rate coefficients rec- values from Klippenstein et al. [396] and Alecu and Truhlar
ommended by Jasper et al. [393] are utilized in the recent [397,398] deviated by a factor of w3.5e4. Aranda et al. [153]
methanol mechanisms proposed by Metcalfe et al. [314] and explained that the discrepancy between the high-level calcula-
Aranda et al. [153]. tions is attributed to the different treatment of anharmoncity. As
will be shown later, a definitive direct experimental measurement
is needed to provide further guidance in reaction mechanism
CH3OH % CH3 þ OH (R1)
development.
CH3OH % 1CH2 þ H2O (R2)
CH3OH þ HO2 % CH2OH þ H2O2 (R8)
CH3OH % CH2OH þ H (R3)
CH3OH þ HO2 % CH3O þ H2O2 (R9)
Hydrogen abstraction from methanol by H radicals have been
studied experimentally prior to 2001, as reviewed in detail by The hydroxymethyl (CH2OH) and methoxy (CH3O) produced
Baulch et al. [352] (Table 5). Recent experimental studies have not by hydrogen abstraction from methanol are primarily consumed
been found in the literature, but a detailed high-level direct dy- via thermal decomposition (i.e., b-scission reactions), leading to
namic variational transition state theory calculation has been per- the products of both reactions being formaldehyde and H radi-
formed by Meana-Pañeda, Truhlar, and Fernández-Ramos [394]. cals. The hydroxymethyl radical can also react with molecular
The authors present good agreement of their calculated total rate oxygen to produce formaldehyde and H radicals. The methoxy
coefficient ((R4) þ (R5)) with the available experimental data. radical is primarily consumed via thermal decomposition (R10).
Furthermore, their calculations indicate a variation in the branch- Rasmussen et al. [147] reviewed rate coefficients for (R10), and
ing ratio with temperature, wherein that (R4) contributes to 100% the preferred rate coefficient is based on the work of Hippler et al.
at room temperature and w75% at 2500 K. [399]. Baulch et al. [352] reviewed the various rate coefficients
proposed for R11, including those by Bowman [123], Tsuboi and
Hashimoto [128], Cribb et al. [135,136], Greenhill and Ogrady
CH3OH þ H % CH2OH þ H2 (R4)
[362], and Held and Dryer [312]. Baulch et al. [352] noted large
discrepancies in the aforementioned rate coefficient recommen-
CH3OH þ H % CH3O þ H2 (R5)
dations. No detailed study of the pressure dependence of this
reaction has been made, even though Greenhill and Ogrady [362]
Hydrogen abstraction by OH radicals is the primary consump-
showed that fall-off effects are important. Baulch et al. [352]
tion pathway for methanol in lean and stoichiometric environ-
recommended using the rate coefficient for (R10) proposed by
ments. Baulch et al. [352] reviewed the large body of experimental
Held and Dryer [312], and this is the one most commonly used in
rate coefficient measurements prior to 2001. In 2003 and 2005,
recent methanol combustion models [147,314]. Baulch et al. [352]
these reactions were studied at atmospheric conditions (<400 K)
also reviewed the proposed rate coefficients for oxidation of
by Jiménez et al. [353] and Dillon et al. [354]. Srinivasan et al.
hydroxymethyl by O2 (R12), and they noted an excellent agree-
[355] recently conducted high-accuracy experiments at high
ment across experimental studies at low temperatures. The high
temperatures. The rate coefficients for (R6) þ (R7) and its
temperature rate coefficient extrapolation from the work of
branching ratios were studied theoretically by Xu and Lin [395]
Grotheer et al. [400] (at 298e684 K) shows good agreement with
using high-level computational theory. Their study noted the
flame studies by Vandooren and van Tiggelen [127]. Baulch et al.
importance of characterizing pre-reaction complexes in this class
[352] recommended a rate coefficient for this reaction by fitting
of reactions. Their proposed rate coefficient in the temperature
the mean of the low-temperature data with the high-temperature
range of 200e3000 K is in good agreement with the available
measurements, resulting in a rate coefficient valid in the range
experimental data. Their calculations indicate a branching ratio
298e1200 K.
that is largely independent of temperature; (R6) contributes to
96e89% of the total rate coefficient across the entire temperature
range, which is in agreement with the recent combustion exper- CH3O % CH2O þ H (R10)
iments of Rasmussen et al. [147]
CH2OH % CH2O þ H (R11)
CH3OH þ OH % CH2OH þ H2O (R6)
CH2OH þ O2 % CH2O þ HO2 (R12)
CH3OH þ OH % CH3O þ H2O (R7)

Hydrogen abstraction from methanol by HO2 has never been The methanol mechanism utilized in this study is that devel-
measured experimentally. Metcalfe et al. [314] utilized brute-force oped recently by Metcalfe et al. [314], namely AramcoMech 1.3.
sensitivity analysis to show that high pressure (30 bar) shock tube This mechanism is chosen due to its comprehensive development
ignition delay times are sensitive to the rate coefficient for (R8). In and testing for methanol, as well as for hydrogen [323], hydrogen/
studies of RCM, Kumar and Sung [151] and Aranda et al. [153] also CO mixtures [323], methane, and formaldehyde, which are all
highlighted the importance of this reaction’s rate coefficient for relevant to methanol combustion. In their paper [314], the au-
accurately simulating RCM ignition delay measurements. The thors present methanol model comparisons against shock tube
recent theoretical investigations include those by Altarawneh et al. ignition delay data from Bowman [123] (>1500 K, 1e11 bar,
[335], Klippenstein et al. [396], and Alecu and Truhlar [397,398]. f ¼ 0.75e6.0) and from Fieweger et al. [142] (750e1250 K, 13e
Metcalfe et al. [314] compared the rate coefficients determined in 40 bar), flow reactor speciation data from Norton and Dryer
various studies, including estimations based on analogies and/or [132,133] and Held and Dryer [140], and premixed laminar flame
comparison with experimental combustion data and theoretically speed data from Veloo et al. [110] and Vancoillie et al. [152] (1 bar,
calculated values. From 900 to 1000 K, the theoretically calculated f ¼ 0.7e1.5).
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 71

Atmospheric pressure methanol/air premixed laminar flame 1.0E+00


13 bar ST
speeds were recently measured by Veloo et al. [110] and Vancoillie
var V
et al. [152] near 343 K. The former utilized the counterflow twin
var V (CH3OH+HO2)
premixed flame technique, whereas the latter employed the heat 30 bar RCM
flux burner method. Egolfopoulos et al. [137] also utilized the 1.0E-01
var V
counterflow twin flame technique to measure methanol/air flame var V (CH3OH+HO2)

Ignition Delay Time (s)


speeds near 340 K, although they used a linear extrapolation to 40 bar ST
account for flame stretch, whereas Veloo et al. [110] utilized a more var V
1.0E-02
accurate non-linear extrapolation to zero stretch. Fig. 19 presents const V
the experimental data obtained in the aforementioned studies. The
recent data sets from Veloo et al. [110] and Vancoillie et al. [152]
agree well for lean and stoichiometric mixtures, but rich mixtures 1.0E-03
are notably faster in the latter data set. The model from Metcalfe
et al. agrees with both data sets under lean conditions and partic-
ularly concurs with the data of Vancoillie et al. [152] on the rich
side, albeit there is only a single measurement at this condition. 1.0E-04
Additional experiments for methanol/air mixtures under rich
conditions are needed to better validate the chemical kinetic
mechanisms.
1.0E-05
The homogenous ignition delay of methanol has been well 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
studied under Ar dilution at high temperatures above 1200 K
1000 K / T
(refer to Table 3). Kumar and Sung conducted RCM measurements
at elevated pressures and intermediate temperatures of greater Fig. 20. Methanol ignition delay times at stoichiometric conditions. Symbols are
relevance to internal combustion engines (7e30 bar, 850e1100 K, experimental data from Fieweger et al. [142] (shock tube, methanol/O2/N2, 13 and
40 bar) and from Kumar and Sung [151] (RCM, methanol/O2/Ar, 30 bar). Open symbols
f ¼ 0.25e2.0), again under Ar dilution (O2/Ar : 1/3.76). Their RCM represent shock tube measurements affected by deflagration. Lines are simulations
data [151] in Fig. 20 displays methanol’s Arrhenius-type mono- under various conditions (see text).
tonic decrease in the ignition delay time with increasing temper-
ature. Simulations are also presented using the chemical kinetic
model of Metcalfe et al. [314] (dashed gray line) and employing a
volume-varying history to account for heat losses after the end of 13e40 bar, f ¼ 1.0). Fig. 20 presents their experimental results
compression [104,106,401,402]. The model is able to simulate the along with simulations using the chemical kinetic model of Met-
qualitative Arrhenius behavior that is observed experimentally, calfe et al. [314]. Methanol/air mixtures display long ignition delay
but the simulated ignition delay times are quantitatively lower times at high pressures and intermediate temperatures (i.e., 10e
than those measured experimentally. Rate coefficients responsible 100 ms), which agrees with its high RON value (i.e., 112). Constant
for the discrepancy are discussed later. Fieweger et al. [142] volume simulations (Fig. 20 thin dashed blue line) are presented at
measured shock tube ignition delay times for methanol/air and 40 bar, and the model displays the linear trend of decreasing
conditions comparable to SI engine conditions (i.e., 750e1250 K, ignition delay time with decreasing temperature. The data suggest
significantly shorter ignition delay times at lower temperatures,
but Fieweger et al. [142] attributed this discrepancy to the exis-
tence of a pronounced deflagration phase that causes the post-
compression pressure to double prior to the main ignition event.
The result is up to an order of magnitude decrease in ignition
delay times for experiments affected by deflagration (e.g., open
symbols in Fig. 20). Increasing pressure during the ignition delay
time period is not ideal, and such phenomena can be modeled
using volume-varying simulations, as explained by Chaos and
Dryer [102] or pressure-varying simulations as explained by Zhu
et al. [240]. When the former methodology (denoted as “vol hist
sim” in Fig. 20) is implemented using a 10%/ms pressure rise rate
(Fig. 20, thick dashed blue line and solid red line), the model
simulations are in much better agreement with the experimental
data.
The oxidation of methanol in stirred and flow reactors has
been studied to measure the global reactivity and to identify
intermediate and product species distributions. A list of pub-
lished studies is presented in Table 3, with the most notable ones
being those at elevated pressures by Held and Dryer [312], Dayma
et al. [146], and Aranda et al. [153]. Metcalfe et al. [314] presented
comprehensive comparisons of their methanol model against the
data of Held and Dryer [312]. Here the model is compared with
JSR data at 10 bar [146] because the same facility and experi-
mental methodology has also been used to study many other
alcohol fuels. The data in Fig. 21 indicate that methanol is
Fig. 19. Laminar flame speeds for methanol/air mixtures at 1 bar and 343 K. Symbols
are experimental measurements by Veloo et al. [110], Vancoillie et al. [152], and unreactive below 800 K and no intermediate or product species
Egolfopoulos et al. [137] and the line is a simulation. are measured. Above 800 K, methanol reacts to produce
72 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Fig. 21. Jet-stirred reactor speciation profiles for methanol oxidation at 10 bar, f ¼ 0.6,
s ¼ 1 s, 8000 ppm CH3OH, 20000 ppm O2, 800 ppm H2O, and balance N2. Symbols are
experimental data from Dayma et al. [146]. Lines are simulations.

Fig. 22. Reaction path diagram for methanol combustion.


formaldehyde (CH2O), which has a peak concentration near
Figure adapted from Ref. [153].
825 K, and is then oxidized to form CO and subsequently CO2. The
model of Metcalfe et al. [314] qualitatively agrees with the con-
sumption of methanol, the production and consumption of and 21. Adopting the recent theoretically calculated rate co-
formaldehyde, and the generation of major product species (CO, efficients from Alecu and Truhlar [397,398] slows down ignition
CO2, and H2O). However, the simulated reactivity is shifted to delay times under RCM conditions and simulations are in better
lower temperatures (w50 K) than measured experimentally. As agreement with the experimental data. Their slower rate coeffi-
will be discussed later, this is attributed to the selected rate co- cient also slows the reactivity of methanol in the JSR bringing it into
efficient for CH3OH þ HO2 % CH2OH þ H2O2. better agreement with the experimental profile and also improving
The general reaction scheme for methanol combustion in pre- formaldehyde agreement. However, the proposed rate coefficient
mixed flames, batch reactors, and stirred and flow reactors is pre- also slows ST ignition delay times, resulting in an under agreement
sented in Fig. 22. Methanol consumption is initiated by hydrogen of the data from Fieweger et al. [142], which is consistent with the
abstraction by OH, HO2, and H radicals. As mentioned earlier, the findings of Metcalfe et al. [314] when using the aforementioned
branching ratio of hydrogen abstraction by OH strongly favors the rate coefficient to simulate ST data from Noorani et al. [150]. The
formation of CH2OH with CH3O being a minor channel. Both these various combustion modeling studies [147,153,313,314] and theo-
radical intermediates share their fate in the formation of formal- retical calculations [335,396,398] have not yet converged upon a
dehyde, and once formed this species controls the reactivity of the
system. Hydrogen abstraction from formaldehyde yields the HCO
radical (formyl), eventually leading to the formation of CO and CO2
via the sequence CH2O / HCO / CO / CO2.
Both Veloo et al. [110] and Metcalfe et al. [314] have shown that
the laminar flame speed of methanol displays a strong positive
sensitivity to the ubiquitous radical chain branching
H þ O2 % O þ OH and the reaction HCO(þM) % H þ CO(þM). The
latter reaction produces the H radicals needed to drive the former
reaction. The reaction of HCO þ O2 % CO þ HO2 has been shown to
inhibit laminar flame speeds because it competes directly with the
HCO decomposition channel and traps H-atoms in the hydro-
peroxyl radical.
Fig. 23 from Kumar and Sung [151] presents a sensitivity analysis
on ignition delay times measured in the RCM at 30 bar, 905 K, and
under stoichiometric conditions, which is one of the points plotted
in Fig. 20. The ignition delay is strongly promoted by the reactions
CH3OH þ HO2 % CH2OH þ H2O2 and H2O2(þM) % OH þ OH(þM).
The former reaction produces the hydrogen peroxide (H2O2) that
Fig. 23. Sensitivity figures of merit obtained by applying Morris analysis on RCM
decomposes to form reactive OH radicals in the latter reaction. The ignition experiments conducted at 30 bar, 905 K, under stoichiometric conditions, on
importance of the reaction CH3OH þ HO2 % CH2OH þ H2O2 under methanol/O2/Ar mixtures.
combustion-relevant conditions is further highlighted in Figs. 20 Reprinted from Kumar and Sung [151] with permission from John Wiley and Sons.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 73

agreeable rate coefficient for CH3OH þ HO2 % CH2OH þ H2O2, and dependence of the branching fraction is smaller than previous
thus a direct experimental measurement is greatly needed. theoretical calculations suggest.

C2H5OH % C2H4 þ H2O (R13)


4.6. The combustion chemistry of ethanol
C2H5OH % CH3 þ CH2OH (R14)
Ethanol is the most widely used alcohol fuel in transportation
applications. Its oxidation mechanism is also an important sub- C2H5OH % C2H5 þ OH (R15)
component of ethane and propane combustion models. Detailed
chemical kinetic models for ethanol have been presented and H-atom abstraction reactions from ethanol proceed from the a,
tested against experimental data. Notable models include the first b, or hydroxyl sites. The total rate coefficient for H-atom abstraction
by Natarajan and Bhaskaran in 1981 [126]; the models by Egolfo- by H radicals ((R16) þ (R17) þ (R18)) was first studied by Aders and
poulos et al. [159] and Marinov [310] in the 1990s that were Wagner in 1973 [408] in the temperature range of 295e700 K. The
compared against data from shock tube ignition delay, laminar only other direct experimental measurement was recently per-
flame speeds, and chemical species concentration measurements in formed by Sivaramakrishnan et al. [366] at temperatures of 1054e
a jet-stirred reactor; and the comprehensively developed model by 1359 K, wherein the total rate coefficient was measured and indi-
Frassoldati et al. [403] for ethanol and ethanolegasoline mixtures. vidual rate coefficients were calculated from theory. The theoreti-
Other recent ethanol oxidation models that are widely tested (i.e., cally calculated rate coefficients from Sivaramakrishnan et al. [366]
shock tube, jet-stirred reactor, flow reactor, flame speed, and flame and Park et al. [409] are in excellent agreement with the afore-
speciation) include AramcoMech presented by Metcalfe et al. [314] mentioned direct measurements. The branching ratios are also in
and Mittal et al. [187] and the ethanol model by Leplat et al. [177]. A agreement, with (R17) being the dominant channel (w99% at 500 K
number of other models have also been presented for ethanol and w64% at 2000 K), (R18) contributing a maximum of 38% at
oxidation, but they tend to present comparisons against a limited 2500 K, and (R16) being negligible (7% at 2500 K) [366].
set of experimental data [110,160,169,173,176,179,181,184,404].
The unimolecular thermal decomposition of ethanol proceeds C2H5OH þ H % C2H5O þ H2 (R16)
primarily via a dehydration reaction (R13), CeC bond scission (R14),
or CeO bond scission (R15). Baulch et al. [352] reviewed the limited C2H5OH þ H % CH3CHOH þ H2 (R17)
number of related experimental studies (e.g., the direct measure-
ment of (R13) by Herzler et al. [405]), but noted that none provided C2H5OH þ H % CH2CH2OH þ H2 (R18)
detailed temperature and pressure-dependent rate coefficients.
More recently, Li et al. [370] measured the rate coefficient of (R13) OH radicals are the most important abstractors of hydrogen
by monitoring ethanol decay and ethylene production in a pyrolysis atoms from ethanol. Table 5 lists experimental measurements for
flow reactor with online FTIR sampling. Their experimentally the total rate coefficient ((R19)-(R21)) across a range of tempera-
determined rate coefficient agreed with that of Herzler et al. [405]. tures. Early works in the 1980s [361,362,376] studied this reaction at
Tsang [406] conducted a master-equation analysis, concluding that low temperatures of relevance to atmospheric chemistry, except for
R13 is the main reaction with CeC and CeO bond scissions being the study of Hess and Tully, who conducted measurements up to
secondary processes, and assigned a rate in creditable agreement 750 K [364]. Bott and Cohen later performed higher temperature
with the aforementioned experimental measurements [370,405]. measurements (1200 K) [365] using a shock tube with laser ab-
Park et al. [407] employed ab initio molecular orbital theory and sorption spectroscopy. In the past decade, the rate coefficient was
RiceeRamspergereKasseleMarcus (RRKM) calculations to deter- again measured by Jiménez et al. [353] and Dillon et al. [354] in at-
mine pressure and temperature-dependent rate coefficients and mospheric temperature ranges and by Carr et al. [367] and Sivar-
branching ratios, but their calculated rate for (R13) was 3.2 times amakrishnan et al. [366] at combustion-relevant temperatures. The
higher than that of Tsang [406]. Sivaramakrishnan et al. [366] consensus among experimental studies is that the total rate coeffi-
recently conducted a definitive experimental and theoretical cient is well established from 200 to 2500 K, but the proposed
analysis using a high-temperature shock tube with ab initio variable branching ratios vary greatly between studies. One challenge is that
reaction coordinate transition state theory-based master equation the individual rate coefficient of (R21) cannot be measured by
calculations. The rate coefficient of (R15) was measured directly monitoring OH radical concentrations because the b CH2CH2OH
using OH optical absorption, the total rate coefficient and that of radical quickly decomposes to C2H4 plus OH above 600 K. Carr et al.
(R13) þ (R14) was measured by H-atom ARAS detection, and the [367] utilized isotopic labeling to measure individual rate co-
difference gave the rate coefficient of (R13). They highlighted the efficients of (R21) and (R22) below 550 K, but these temperatures
sources of disagreement with previous theoretical calculations, were not directly relevant to combustion. Thus, theoretical calcu-
noting that Park et al. [407] did not correctly account for the bar- lations and/or estimations are necessary to better determine the
rierless potential of (R13), while Li et al. [370] artificially lowered individual rate coefficients. Such theoretical calculations have been
their transition state frequencies to match experimental data. In performed by Xu and Lin [395], Zheng and Truhlar [410], and
addition, the variations in pressure fall-off treatment from the use Sivaramakrishnan et al. [366]. In general, all these studies present
of different energy transfer models led to discrepancies in theo- total rate coefficients that can reproduce the aforementioned
retical rate coefficient calculations. The branching ratios from experimentally measured data. However, Sivaramakrishnan et al.
Sivaramakrishnan et al. [366] favor the dehydration reaction (R13) [366] had to adjust their theoretical calculations to obtain good
below 1400 K (ranging from 90 to 70% from 500 to 1400K), while agreement with data, whereas Zheng and Truhlar [410] noted that
CeC bond scission (R14) becomes increasingly important at high their lower level potential energy surface calculations (MS-VTST
temperature (w45% at 2000 K). The CeO bond scission (R15) in with M08-SO/6-31 þ G(d,p)) agreed with experimental data better
ethanol is negligible because its bond strength is significantly than did their state-of-the-art multi-path variational transition state
higher than that of a CeC bond. A recent shock tube study by Wu theory calculations (MP-VTST with M08-HX/6-31 þ G(d,p)).
et al. [371] on ethanol decomposition using H-atom ARAS detection Table 6 presents branching ratios determined in several recent
are in good agreement with previous works; however, the pressure studies. The theoretical calculations of Xu and Lin [395], Zheng and
74 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Table 6
Branching ratios for hydrogen abstraction by OH radicals from ethanol’s a, b, and OH sites at 900 K.

Sivaramakrishnan et al. [366] Xu and Lin [395] Zheng and Truhlar [410] Carr et al. [367] Tran et al. [184]

% of total rate a 93 83 74 39 79
b 4 13 15 10 21
OH 3 4 11 51 3

Truhlar [410] (MP-VTST calculations), and Sivaramakrishnan et al. n-butanol by hydroperoxyl radicals showed that hydrogen bonds in
[366] (adjusted theoretical calculations), and the estimate from the transition state lead to ring structures. These were found to be
Tran et al. [184] (EvansePolanyi correlation with ethane) all suggest important for the b and g sites of n-butanol, resulting in lower
that abstraction from the a site (R20) is dominant followed by the b energy barriers and higher entropy loss (i.e., lower activation en-
channel (R21). The rate coefficients proposed by Carr et al. [367] ergies and lower A-factors). Sarathy et al. [260] found that the rate
suggest that (R19) is the dominant channel, which is unexpected coefficients determined by Zhou et al. [326] for n-butanol were
given the high BDE of the OeH bond, and in disagreement with the suitable for combustion modeling; however, an increase in the a
other studies. The branching ratio of (R20) varies by about 20% channel rate by a factor of 2.5 (within the uncertainty specified by
between Zheng and Truhlar [410] and Sivaramakrishnan et al. Zhou et al. [326]) was needed to obtain better agreement with the
[366], wherein the former places more importance on (R19) and results of intermediate temperature shock tube ignition delay ex-
(R21). Zheng and Truhlar [410] have identified inaccuracies in periments. Metcalfe et al. [314] and Mittal et al. [187] expected that
previous studies [366,395], such as unexpectedly high saddle point such hydrogen bonding effects would also be important in ethanol.
energies for transition states with hydrogen bonded structures in Mittal et al. [187] assigned (R23) a rate coefficient 1.75 times greater
(R21), which explains why Sivaramakrishnan et al.’s [366] than the a channel rate recommended by Zhou et al. [326] for n-
branching ratio for (R21) is lower than others. The MP-VTST cal- butanol. Their rate coefficient for abstraction from the hydroxyl
culations by Zheng and Truhlar [410] are the state-of-the-art in moiety (i.e., (R22)) was that calculated by Zhou et al. [326] for the
computational chemistry. It seems reasonable that their branching analogous reaction in n-butanol. Abstraction from the b carbon was
ratios would therefore be the most accurate given the high-level of estimated as the average of Zhou et al.’s [326] calculated rate co-
theory employed. However, we recall that their total rate coefficient efficients from the d and b channels in n-butanol.
could not accurately reproduce experimental data. They [410]
concluded that advances in computational chemistry are needed C2H5OH þ HO2 % C2H5O þ H2O2 (R22)
to make rate calculations more predictive, especially for reactions
with stabilized pre-reaction complexes and hydrogen bonding in C2H5OH þ HO2 % CH3CHOH þ H2O2 (R23)
the transition states.
C2H5OH þ HO2 % CH2CH2OH þ H2O2 (R24)
C2H5OH þ OH % C2H5O þ H2O (R19)
At high temperatures, the ethoxy (C2H5O), a-hydroxyethyl
C2H5OH þ OH % CH3CHOH þ H2O (R20) (CH3CHOH), and b-hydroxyethyl (CH2CH2OH) radicals produced by
hydrogen abstraction from ethanol are consumed via thermal
C2H5OH þ OH % CH2CH2OH þ H2O (R21) decomposition (i.e., b-scission reactions) and unimolecular

Hydrogen abstraction from ethanol by HO2 has not yet been


measured experimentally or calculated using theoretical chemistry.
As shown previously for methanol and later for ethanol, ignition
delay times of ethanol/air mixtures are sensitive to the total reac-
tion rate and its branching ratios. One would expect the abstraction
to favor the a site given its low BDE and the abstraction from the
hydroxyl moiety to be negligible. Recent chemical kinetic models
vary widely in their choice of rate coefficients for (R22), (R23), and
(R24). Several proposed rate coefficients for (R23) (a) and (R24) (b)
are shown in Fig. 24. All studies suggest that abstraction from the a
site is dominant at all temperatures, but the temperature depen-
dence varies across studies. Estimates of (R24) by Marinov [310]and
Tran et al. [184] agree with each other at 1000 K, whereas that of
Mittal et al. [187]is nearly a factor of 3 higher. Leplat et al. [177] and
Lee et al. [181] used the proposed rate coefficient from Marinov
[310]. He assumed that (R23) had a rate coefficient similar to
methanol’s (R8) (CH3OH þ HO2 % CH2OH þ H2O2) and that the rate
coefficient of (R24) was based on the derived rate expression of
(R23). Marinov [310] assigned large error bars to his proposed rate
coefficients and stated that they were highly uncertain. Recently,
Tran et al. [184] proposed rate coefficients for (R23) and (R24) using
EvanePolanyi correlations proposed by Dean and Bozzelli [411]
with ethane as a reference. This methodology accounts for varia-
tions in BDEs between ethanol and ethane, but not for secondary
Fig. 24. Comparison of rate coefficients for C2H5OH þ HO2 from a and b sites. Rate
effects such as hydrogen bonds formed in transition states. A recent coefficients are plotted from the studies of Mittal et al. [187], Marinov [310], and Tran
theoretical study by Zhou et al. [326] on hydrogen abstraction from et al. [184].
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 75

isomerization reactions. Xu et al. [412] recently conducted a theo-


retical study on the decomposition and isomerization of C2H5O
radicals. Senosiain et al. [413] also performed high-level quantum
chemistry calculations for all the low-energy pathways of the re-
action of ethylene with OH radicals. These included all C2H5O iso-
mers following OH addition to the double bond in ethylene and
subsequent isomerization reactions. Both studies noted that very
few direct measurements have been made at combustion-relevant
temperatures. The ethoxy radical (C2H5O) undergoes decomposi-
tion either by CeC bond scission (R25) to formaldehyde and methyl
radical or CeH bond scission (R26) to acetaldehyde and H radical,
with isomerization reactions being negligible. At 1 bar and 1000 K,
the branching ratio of R25:R26 is approximately 73:27 [412]. The a-
hydroxyethyl (CH3CHOH) also favors thermal decomposition at
combustion-relevant conditions. The OeH bond scission forming
acetaldehyde (R27) is in competition with the CeH bond scission
leading to ethenol (R28) [413]. Both channels are thus important,
but the path leading to acetaldehyde is favored according to cal-
culations by Dames [414]. CeC bond scission to ethylene and OH
radical (R30) accounts for about 90% of the destruction of b-
hydroxyethyl (CH2CH2OH), with 10% going to form ethenol (R31)
[412].

C2H5O % CH2O þ CH3 (R25)

C2H5O % CH3CHO þ H (R26)

CH3CHOH % CH3CHO þ H (R27)

CH3CHOH % C2H3OH þ H (R28) Fig. 25. Stationary points on the (a) CH3CHOH þ O2 and on the (b) CH2CH2OH þ O2
potential energy surfaces.
Reprinted from Zádor et al. [339] with permission from Elsevier.
CH2CH2OH % C2H4 þ OH (R29)

CH2CH2OH % C2H3OH þ H (R30)


surface (PES) for the b-hydroxyethyl þ O2 reactions (Fig. 25) is also
At lower temperatures (below 1000 K) the reactions between a- complicated, but Zádor et al.’s [339] theoretical and experimental
and b-hydroxyethyl and O2 are important because the aforemen- results indicate two pathways as dominant. First, as previously
tioned decomposition channels are slow. Rate coefficients for these discussed, the Waddington pathway is shown to be important,
reactions were studied experimentally at low pressures (<10 mbar) wherein a b-RO2 radical is stabilized and then quickly decomposes
by Miyoshi et al. [415] at room temperature using the laser flash to two formaldehyde and one OH radical (R32). The second
photolysis-photoionization mass spectrometry method, and by important reaction is a concerted elimination reaction leading to
Grotheer et al. [416] in the temperature range of 300e682 K using a the formation of ethenol and HO2 radicals (R33). The branching
discharge flow reactor with a low electron energy mass spec- ratio between (R32) and (R33) is still uncertain due to the multi-
trometer. Zádor et al. [339] recently conducted a theoretical anal- reference character of the transition states
ysis and experimental product study of the reaction of
hydroxyethyl radicals with O2. Their experiments studied the pri- CH3CHOH þ O2 % CH3CHO þ HO2 (R31)
mary products of the reaction (not rate coefficient measurements)
initiated by an excimer laser in a heated flow reactor coupled with CH2CH2OH þ O2 % OOCH2CH2OH % 2 CH2O þ OH (R32)
tunable synchrotron photoionization mass spectrometry. Pressure-
dependent rate coefficients and branching ratios were computed CH3CHOH þ O2 % C2H3OH þ HO2 (R33)
(and not measured) in the temperature range of 255e1000 K using
high-level ab initio potential energy surface calculations and state- The present analysis utilizes the ethanol mechanism by origi-
of-the-art master equation methods. Around the same time as nally presented by Metcalfe et al. [314] and recently modified by
Zádor et al. conducted their work, the reactions of a-hydroxyethyl Mittal et al. [187]. This mechanism is chosen due to its compre-
with O2 were also studied theoretically by da Silva et al. [340]. The hensive development and testing for ethanol combustion. In their
potential energy surfaces of these reactions are complex, as shown papers [187,314], the authors present ethanol model comparisons
in Fig. 25 from Zádor et al. [339], but theoretical studies help to against RCM ignition delay data, (825e985 K, 10e50 bar, and
identify the dominant pathways. Both theoretical studies are in f ¼ 0.3e1.0) shock tube ignition delay data by Heufer and Olivier
excellent agreement regarding the CH3CHOH þ O2 reaction, [174] (down to 800 K, 13e40 bar, f ¼ 1.0); JSR speciation data from
concluding that it proceeds through an activated a-hydroxy-eth- LePlat et al. [177] (890e1250 K, 1 bar, f ¼ 0.25e2) and Dagaut and
ylperoxy adduct that rapidly decomposes to acetaldehyde þ HO2 Togbé [169] (770e1220 K, at 10 bar, f ¼ 0.6e2); flow reactor
(R31). The predominance of (R31) was confirmed experimentally speciation data from Li et al. [161] (1045e1080 K, 1.7e3.0 bar) and
by Zádor et al. [339]. Furthermore, the decomposition reaction Norton and Dryer [133] (1020e1120 K initial, 1 bar, f ¼ 1.18); pre-
occurs with such a low barrier that conventional low-temperature mixed laminar flame speeds from Egolfopoulos et al. [159] (1 bar,
chain branching pathways are unimportant. The potential energy f ¼ 0.6e1.8), van Lipzig et al. [178] (1 bar, f ¼ 0.6e1.5), and Gülder
76 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

[129] (1e8 bar, f ¼ 0.7e1.4); and premixed laminar flame speciation


data from Leplat et al. [177] (500e1800 K, 50 mbar, f ¼ 0.75e1.25).
The laminar flame speeds of ethanol/air mixtures have been
measured in a number of studies, as noted in Table 3, and there is
reasonably good agreement across the various measurements ob-
tained at atmospheric pressure and a range of unburned gas tem-
peratures. Metcalfe et al. [314] and Mittal et al. [187] have shown
that their ethanol model can accurately reproduce laminar flame
speeds data at 298 K [129,159,178], 338 K [178], 363 K [159,178],
428 K [159], and 453 K [159]. Not presented in their paper is the
performance of the model at elevated pressures. Fig. 26 presents
the experimental data from Bradley et al. [172], which covers
pressures from 1 bar up to 14 bar and a range of equivalence ratios.
Their maximum equivalence ratio achievable at higher pressures
was constrained by the safe operating pressure limit of their
combustion vessel. The data indicates a decrease in flame speed as
pressure increases, a qualitative trend that is accurately reproduced
by the model. The quantitative agreement between the simulations
and measurements are in need of improvement. At 1 bar, the pre-
sent model simulates higher flame speeds than the measured data,
which is a discrepancy not observed when comparing with data
from other facilities [159,314]. At 2 bar, the model simulates higher
flame speeds than the data by up to 10 cm/s, whereas at higher
pressures the model simulations are generally within 5 cm/s of the Fig. 27. Ignition delay times of ethanol/air mixtures at stoichiometric conditions.
experimental data. The sensitive reactions controlling the model Symbols are experimental data from Cancino et al. [173] (shock tube at 10, 30, and
50 bar) and Lee et al. [181] (RCM at 40 bar and ST at 80 bar) and lines are simulations.
simulations are small molecular reactions [314] and not specific to Shock tube simulations account for pre-ignition pressure rises.
ethanol chemistry. It should be noted that the ethanol model pro-
posed by Leplat et al. [177] provides much better quantitative ethanol/air ignition delay times at a range of pressures using shock
agreement with the data from Ref. [172] at all pressures. tubes [173,181] and RCMs [181] along with chemical kinetic
The homogenous ignition delay of ethanol/air has been studied modeling simulations that account for shock tube pressure rises
at a range of pressures and temperatures of relevance to internal prior to ignition. The RCM measurements at 30 bar exhibit Arrhe-
combustion engines using both shock tubes and RCMs. A recent nius behavior below 1000 K, which is extended to higher temper-
shock tube ignition study by Lee et al. [181] utilized schlieren im- atures in the shock tube measurements. The discrepancy between
aging to show that pre-ignition reactions cause significant pressure measured ignition delay times in the temperature region covering
rises prior to the main ignition event, and these pressure rises must both RCM and shock tube experiments is attributed to the afore-
be well characterized for comparison against chemical kinetic mentioned pre-ignition pressure rise phenomenon in shock tubes.
modeling simulations. Fig. 27 presents experimentally measured The model is able to well reproduce both RCM and shock tube when
facility effects are included, along with the experimentally
observed decreasing ignition delay times with increasing pressure.
The lack of NTC behavior in ethanol is consistent with its high
research octane number. The combustion chemistry contributing to
these features is discussed later.
Speciation measurements during oxidation of ethanol have
been studied extensively in flow reactors, stirred reactors, non-
premixed flames, and premixed flames using a variety of diag-
nostic techniques (refer to Table 3). Here, the ethanol kinetic model
is compared with JSR data at 10 bar from Dagaut and Togbé [169]
and low-pressure premixed flame speciation measurements using
MBMS by Leplat et al. [177]. Only selected species profiles are
presented in Fig. 28 to highlight some important features of ethanol
oxidation and the model’s performance. The low-pressure pre-
mixed flames are burner stabilized, allowing the experimentally
measured temperature profile to be used as a boundary condition
for the simulation. In both the JSR and low-pressure flame, ethanol
is oxidized to form acetaldehyde and ethylene, which are primary
b-scission products. The model accurately reproduces the qualita-
tive formation of these intermediates in the JSR and their con-
sumption, but ethanol reacts at lower temperatures to form
acetaldehyde in the model than in the experiments. Ethylene
concentrations are correctly simulated in both the JSR and with the
low-pressure premixed flame. Later, we highlight some of the
important reactions controlling these products, namely hydrogen
Fig. 26. Laminar flame speeds for ethanol/air mixtures at 1e14 bar pressures and
abstractions by OH and HO2 radicals. Ethenol (C2H3OH) is an
358 K. Symbols are experimental measurements by Bradley et al. [172] and lines are interesting species that was not measured in the JSR or ethanol
simulations. low-pressure flames due to experimental limitations; however, Xu
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 77

Fig. 28. (Left) Jet-stirred reactor speciation profiles for ethanol oxidation [169] at 10 bar, f ¼ 1.0, s ¼ 0.7 s, 2000 ppm C2H5OH, 6000 ppm O2, and balance N2. (Right) Low-pressure
flame speciation profiles [177] for ethanol/O2/Ar at 50 mbar and f ¼ 0.75. Lines are simulations. Simulated flame profiles are shifted by 0.5 mm away from the burner surface.

et al. [179] recently acquired speciation measurements using mo- OH favors the formation of CH3CH2OH with CH3O and CH2CH2OH
lecular beam flame-sampling photoionization mass spectrometry. being less important channels. At high temperatures, these radicals
Ketene (CH2CO) has also been measured in low-pressure flames undergo b-scission wherein CeC and CeO bonds are favored,
[177,179]. Simulations of ethenol and ketene under the low- leading to the formation of stable intermediates, including
pressure premixed flame conditions of Xu et al. [179] were not ethylene, acetaldehyde, and formaldehyde. Highlighted in the
possible because their experimentally measured temperature figure is also the b-scission of a CeH bond in CH3CHOH leading to
profiles are not in agreement with those used for their simulations, the formation of ethenol, a pathway responsible for producing this
and no discussion of this discrepancy is offered in their paper. compound in premixed ethanol/O2 flames (Fig. 28). The acetalde-
Therefore, simulations of ethenol and ketene are presented under hyde intermediate is primarily consumed via hydrogen abstraction
the flame conditions of Leplat et al. and presented in Fig. 28. The from the aldehydic carbon; the subsequent radical then de-
chemistry contributing to the formation of these intermediates is composes to form carbon monoxide. A minor pathway of acetal-
discussed later. dehyde consumption proceeds via abstraction from the methyl
The general reaction scheme for ethanol combustion in pre- group leading to ketene production. This is the primary route to
mixed flames, batch reactors, and stirred and flow reactors is pre- ketene and explains why its maximum concentration in flames
sented in Fig. 29. Ethanol consumption is initiated by hydrogen (Fig. 28) occurs after the acetaldehyde maximum. As shown earlier,
abstraction by radicals (denoted as R), such as OH, HO2, and H. As ethylene concentrations are correctly simulated in the JSR and in
mentioned earlier, the branching ratio of hydrogen abstraction by premixed flames (Fig. 28). This agreement is due to the branching

Fig. 29. Reaction path diagram for ethanol combustion. R denotes a radical, such as OH, HO2, and H. Low-temperature pathways are shown in green. (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)
78 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

ratio of hydrogen abstraction by OH. The model by Metcalfe et al. temperature reactivity, and this phenomenon will be shown to be
[314] utilizes rate coefficients from Sivaramakrishnan et al. [366], important in larger alcohols as well.
which strongly favors the formation of a-hydroxyethyl (CH3CH2OH) Fig. 30 from Ref. [187] presents a ‘brute force’ sensitivity analysis
(Table 6), while Mittal et al. [187] reduced the branching ratio to a- on shock tube ethanol/air ignition delay times at 30 bar, 1050 K, and
hydroxyethyl by 25% and increased the branching to b-hydrox- stoichiometric conditions. These results are consistent with the
yethyl radical by 25%. Calculations by Zheng and Truhlar [314,410] sensitivity analysis conducted in other studies in shock tubes,
give lower weighting to the formation of b-hydroxyethyl RCMs, and JSRs [177,181,279]. The ‘brute force” sensitivity analysis
(CH2CH2OH). Fig. 28 presents simulations using the C2H5OH þ OH quantifies the change in ignition delay time from a doubling and
rate coefficients proposed by Tran et al. [184] that have total rate halving of each reaction rate coefficient. Similar to those previously
coefficients in agreement with experimental results and branching shown for methanol/air mixtures, the ignition delay is strongly
ratios similar to Zheng and Truhlar [410]. The decreased branching promoted by the hydrogen abstraction by HO2 radicals and H2O2
ratio assigned to the b-site decreases ethylene production and in- decomposition. The importance of the reactions involving
creases acetaldehyde formation. Simulations were also conducted C2H5OH þ HO2 under combustion-relevant conditions is further
to determine the importance of ethenol under the JSR and flame highlighted in Figs. 27 and 28, wherein the rates for abstraction
conditions in Fig. 28. Ethenol was simulated in very low concen- from the hydroxyl, a, and b sites are replaced by those proposed by
trations in the JSR, but simulated low-pressure flame concentra- Tran et al. [184]. Recall from Fig. 24 that the total rate coefficient
tions were found to be above 1000 ppm. A reaction path analysis employed by Mittal et al. [187] is approximately a factor of 2 higher
showed that under the 10 bar JSR conditions, the decomposition of than that proposed by Tran et al. [184]. Utilizing the latter rate
CH2CH2OH is primarily to C2H4 þ OH, whereas in the low-pressure coefficient slows the reactivity of ethanol in the JSR and improves
flame, the pathway to C2H3OH þ H equally as important. Thus, the agreement with the experimental data. However, the slower rate
pressure dependence of these reactions alters the branching ratio coefficient worsens the agreement of ignition delay times at 10 bar
of CH2CH2OH decomposition, which is a phenomenon that has not (Fig. 27) and higher pressures. Thus, direct experimental mea-
been studied in great detail to date. surements and/or high-level theoretical calculations are needed to
At low temperatures (i.e., below 1000 K), the a- and b-hydrox- constrain this rate coefficient.
yethyl radicals can react with O2 to form products. As mentioned
previously, the primary product from the a-hydroxyethyl channel is 4.7. The combustion chemistry of propanol isomers
acetaldehyde þ HO2. These reactions control the low-temperature
ignition behavior of ethanol/O2 mixtures, especially at elevated The two isomers of C3 alcohols are n- and iso-propanol. These
pressures. The low energy barrier for the chemically activated a- fuels have received little attention from the combustion chemistry
hydroxyethyl þ O2 channel precludes conventional peroxy stabili- community with few experimental and kinetic modeling studies
zation and isomerization channels that eventually lead to chain available in the literature (refer to Table 3). This lack of combustion
branching processes in alkanes (e.g., ethane, propane, etc.) [340]. data together with a scarcity of fundamental mechanistic and ki-
The b-hydroxyethyl þ O2 channel proceeds through stabilization of netic rate coefficient studies (refer to Table 5) has hindered progress
a b-peroxy (b-RO2) radical that either undergoes a concerted in the development of comprehensive chemical kinetic models for
elimination to form ethenol þ HO2 or a Waddington isomerization the propanol isomers. Frassoldati et al. [195] and Man et al. [200]
and decomposition channel to form two formaldehydes þ an OH have presented kinetic models for the propanol isomers using
radical. The b-channels are of minor importance in combustion primarily analogies with ethanol and the butanol isomers to assign
systems because initial hydrogen abstractions by OH radicals rate coefficients. For the present analysis, we use the sub-
strongly favor the a site at lower temperatures. The absence of an mechanism recently proposed by Man et al. [200], which is
OH radical chain branching route in ethanol explains its lack of low- essentially an updated version the propanol isomers kinetic model

Fig. 30. ‘Brute force’ sensitivity of ignition delay times at 25 bar, three temperatures, and three equivalence ratios.
Reprinted from Mittal et al. [187] with permission from Elsevier.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 79

presented by Johnson et al. [193]. We also added several important nC3H7OH þ R % CH3CHCH2OH þ RH (R36)
low temperature reaction pathways based the reaction classes
explained previously and rate rules derived from the modeling of nC3H7OH þ R % CH2CH2CH2OH þ RH (R37)
butanol isomers [260].
Comparing the bond dissociation energies for n- and iso-prop- iC3H7OH þ R % C3H7O þ RH (R38)
anol with those of other alcohols allow the estimation of rate co-
efficients for unimolecular decomposition reactions. The water iC3H7OH þ R % (CH3)2COH þ RH (R39)
elimination channels are important routes of fuel consumption at
high temperatures, and the rate coefficients for these reactions are iC3H7OH þ R % CH3(CH2)CHOH þ RH (R40)
estimated based on analogies to n-butanol and 2-butanol
[200,260]. Similarly, other CeC bond scissions in the propanol The laminar flame speeds of the propanol isomers in air have
isomers can be assigned rates based on analogous reactions for the been measured in a few studies. Beeckmann et al. [154] noted that
butanol isomers. The rate coefficients of H-atom abstraction re- the n-propanol chemical kinetic model proposed by Johnson et al.
actions from the propanol isomers are not well studied. Abstrac- [193] could not well reproduce laminar flame speeds measure-
tions from n-propanol proceed from the hydroxyl site (R34), a ments by Veloo et al. at 1 bar [198] or their own 10 bar data [154].
(secondary, (R35)), b (secondary, (R36)), or g (primary, (R37)). Ab- Fig. 32 displays experimental [198] and simulated laminar flame
stractions from iso-propanol proceed from the hydroxyl site (R38), speeds for n-propanol/air and iso-propanol/air mixtures at 1 bar.
a (tertiary, (R39)), or two b sites (primary, (R40)). As shown in The present chemical kinetic model well reproduces experimen-
Table 5, there is only one rate coefficient measurement of tally measured values for both fuels across the entire range of
abstraction by OH radical at combustion relevant temperatures, tested conditions. The simulations and experiments indicate that
performed by Dunlop and Tully [372]. Theoretical reaction rate iso-propanol is significantly less reactive than n-propanol, which
studies for hydrogen abstraction from the propanol isomers are indicates that methyl substitution decreases flame speeds. This
lacking. Thus, the site-specific rate coefficients for abstraction re- effect of branching on flame speed is well established in alkane
actions by important radicals (e.g., H, OH, and HO2) must be derived combustion; increased branching decreases formation of reactive
using analogy with reaction rate coefficients established for n- H radicals and increases formation of stable allyl radicals [109,417].
butanol and 2-butanol [200,260]. This approach is tested for the The homogenous ignition delay times of n-propanol/air and iso-
case of iso-propanol þ OH in Fig. 31, wherein experimental mea- propanol/air mixtures have been studied at a limited number of
surements of the total rate coefficients from Dunlop and Tully [372] conditions, primarily at higher temperatures and lower pressures
are compared with estimates based on analogy with 2-butanol [150,193,196,200]. Ignition studies at engine-relevant conditions
[200,260]. The method of estimation using EvansePolyani corre- (e.g., lower temperatures and higher pressures) are not available for
lations [411] is within 25% of the experimentally measured values, testing chemical kinetic models. Fig. 33 presents comparisons of
which indicates the appropriateness of such techniques for esti- kinetic modeling simulations and experimentally measured shock
mating rate coefficients. tube ignition delay times for n-propanol/air and iso-propanol/air
mixtures at pressures ranging from 1.2 to 16 bar, temperatures above
nC3H7OH þ R % C3H7O þ RH (R34) 1100 K [200], and equivalence ratios ranging from lean to rich. The
shock tube measurements display Arrhenius behavior across the
nC3H7OH þ R % CH3CH2CHOH þ RH (R35)

Fig. 31. Comparison of measured (symbols) and estimated (line; see main text for Fig. 32. Laminar flame speeds for n-propanol/air and iso-propanol/air mixtures at
source) total reaction rate coefficients for hydrogen abstraction from iso-propanol by 1 bar and 343 K. Symbols are experimental measurements by Veloo et al. [198] and
OH radical. Measurements are from Dunlop and Tully [372]. lines are simulations.
80 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Fig. 33. Ignition delay times of n-propanol/air (left) and iso-propanol/air (right) mixtures at stoichiometric conditions. Symbols are shock tube ignition delay data from Man et al.
[200] (shock tube at 1.2, 4, and 16 bar) and lines are simulations.

entire range of conditions. iso-Propanol is notably slower to ignite form CO and ethyl radical (C2H5). The b-scission products of fuel
(i.e., less reactive) that n-propanol at the same conditions. These radicals from iso-propanol are propene, ethanol, acetaldehyde, and
features, as well as equivalence ratio effects, are well reproduced by acetone. Subsequent decomposition of propene and acetone both
the kinetic modeling simulations. The present set of tested condi- form resonantly stabilized radicals, which inhibit reactivity.
tions does not allow us to draw conclusions on the lower tempera- Therefore, iso-propanol exhibits lower flame speeds and longer
ture (i.e., below 1000 K) reactivity of the propanol isomers. high temperature ignition delay times that n-propanol because a
The general reaction scheme for n-propanol and iso-propanol larger concentration of stable intermediates are formed. At low
combustion in premixed flames and batch reactors is presented in temperatures (i.e., below 1000 K), the fuel-derived radicals react
Figs. 34 and 35, respectively. The combustion of the propanol iso- with O2 to form a range of products. The primary product from n-
mers is initiated by hydrogen abstraction by radicals (e.g., OH, HO2, propanol’s a-hydroxypropyl channel is propanal þ HO2, while iso-
H, CH3, etc.). As in the case of ethanol, the hydrogen abstraction propanol’s a-hydroxypropyl channel yields acetone þ HO2. For both
reactions favor the formation of a-hydroxypropyl radicals. At high propanol isomers, the b-hydroxypropyl þ O2 channel proceeds
temperatures, the b-scission products of fuel radicals from n- through stabilization of a b-peroxy (b-RO2) radical that undergoes a
propanol are ethylene, propene, ethenol, propanal (i.e., propion- Waddington isomerization and decomposition channel to form one
aldehyde), and formaldehyde. Propanal further decomposes to formaldehyde, one acetaldehyde, and an OH radical.

n-

Fig. 34. Reaction path diagram for n-propanol combustion. R denotes a radical, such as OH, HO2, and H. Low-temperature pathways are shown in green. (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 81

iso-propanol

iso-propoxy

Fig. 35. Reaction path diagram for iso-propanol combustion. R denotes a radical, such as OH, HO2, and H. Low-temperature pathways are shown in green. (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)

4.8. The combustion chemistry of butanol isomers counterflow diffusion flame species profiles, high-temperature
shock tube ignition delay times, doped methane diffusion flame
The combustion chemistry of butanol isomers has been an area species profiles, and n-butanol pyrolysis experiments. Subse-
of intense research activity in the recent six years. The primary quently, the RMG software was utilized by Van Geem et al. [216] to
focus of this work has been on n-butanol because industrialists develop chemical kinetic models for all four butanol isomers. Grana
have targeted it as an alternative fuel for deployment in the fore- et al. [213] also developed a hierarchical chemical kinetic model
seeable future [418]. Nevertheless, the other butanol isomers are with detailed sub-mechanisms for all four butanol isomers. Their n-
also gaining industry attention, and studying all the butanol iso- butanol model was also widely compared against experimental
mers as a family of fuels can elucidate unique structural effects (e.g., speciation data from JSRs and counterflow diffusion flames, pre-
methyl and OH substitution) on the global combustion properties mixed laminar flame speeds, and high-temperature shock tube
and combustion emission profiles of alcohols. ignition delay times. Furthermore, Grana et al. [213] also included
The first chemical kinetic model for n-butanol combustion was models for the other butanol isomers with comaparisons presented
developed by Dagaut and coworkers [208,419] with comparisons for all four isomers against high-temperature shock tube ignition
against JSR speciation data (800e1100 K and 10 bar). Around the delay data [207], iso-butanol species profiles in counterflow diffu-
same time, Moss et al. [207] conducted high-temperature shock sion flames, and tert-butanol oxidation in a flow reactor [133].
tube ignition delay measurements (1200e1800 K and 1e4 bar) and Despite the noteworthy modeling studies prior to 2011, the
proposed a chemical kinetic model for all the four isomers of models did not include the low-temperature peroxy chemistry
butanol using the EXGAS software [207,316]. Later, Sarathy et al. needed to simulate ignition at engine-relevant conditions (e.g., low
[211] extended the comparisons of Dagaut et al.’s butanol model temperatures and high pressures). Again, the limiting factor was
[208,419] to atmospheric pressure measurements of JSR speciation, the lack of experimental data at such conditions to motivate such
premixed laminar flame speed, and counterflow diffusion flame modeling studies. Heufer et al. [219] were the first to conduct shock
speciation. Black et al. [212] also proposed an n-butanol chemical tube ignition delay measurements of n-butanol at low tempera-
kinetic model that was compared against Dagaut and coworker’s tures and high pressures (10e42 bar and 770e1250 K). Weber et al.
JSR data [208] and their own shock tube measurements (1100e [104] later studied the ignition of n-butanol in an RCM at elevated
1800 K, 1e8 bar). The aforementioned n-butanol combustion pressures and low to intermediate temperatures (15 and 30 bar,
studies laid important foundations, but all were limited in their 675e925 K, and f ¼ 0.5, 1.0, and 2.0). Both these studies demon-
targets because of the unavailability of data in various reactive strated that butanol models simulated longer ignition delay times
environments. The first attempts to model a large set of experi- (by more than a factor of 10) at temperatures below w950 K,
mental data for n-butanol were made by Harper et al. [121] and suggesting that important reaction pathways were missing from
Grana et al. [213]. Harper et al. [121] utilized the automatic Reaction the models. Shortly thereafter, Vranckx et al. [223] obtained addi-
Mechanism Generator (RMG) software [420,421] to generate an n- tional n-butanol shock tube ignition data near 80 bar (795 Ke
butanol combustion model capable of reproducing JSR speciation, 1200 K). Their experimental results showed that n-butanol exhibits
82 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

NTC behavior at elevated pressures similar to alkanes, thus moti- hydrogen atoms for water elimination, and not by the position of
vating the researchers to develop a simplified sub-mechanism for the OH group (primary, secondary, or tertiary). Shock tube species
butylperoxy formation capable of simulating their experimental time-history experiments by Stranic et al. [111,112] provide
data. This simplified peroxy mechanism was later revised in a experimental data under pyrolysis conditions that are sensitive to
comprehensive modeling study by Sarathy et al. [260], which unimolecular decomposition rates. A recent modeling study by
included detailed low-temperature reaction pathways for all Vasu and Sarathy [238] showed that the majority of data from
butanol isomers. The model by Sarathy et al. [260] is capable of Stranic et al. could by accurately simulated by adopting the
accurately simulating ignition of n-butanol, iso-butanol, 2-butanol, aforementioned unimolecular decomposition rate constants pro-
and tert-butanol at low temperatures and elevated pressures, as posed by Rosado-Reyes et al.
will be demonstrated later. For the iso-butanol isomer, Merchant et H-atom abstraction reactions from n-butanol proceed from one
al. have published models of comparable accuracy [232,233] that of five sites (a, b, g, d, or hydroxyl). 2-butanol (sC4H9OH) also has
also includes species and reactions leading to formation of aromatic five unique abstraction sites with a hydroxyl site (R41), a site
byproducts. Furthermore, the model developed by Merchant et al. (secondary, (R42)), two unique b sites (a primary, (R43), and sec-
was developed independently yet well predicts a wide range of ondary, (R44)), and a g site (primary, (R45)). iso-Butanol has three
experimental data, and thus demonstrates the ability to formulate a unique abstraction sites (a, b, g, and hydroxyl), whereas tert-
predictive model a priori. butanol has two (three bs and a hydroxyl).
The above background discussion exhibits the data-driven
methodology that led to improved models for low-temperature sC4H9OH þ R % C4H9O þ RH (R41)
combustion of the butanol isomers. As mentioned previously, the
low-temperature ignition behavior of methanol, ethanol, and sC4H9OH þ R % CH3CH2C(CH3)OH þ RH (R42)
propanol isomers can be modeled successfully without the need for
alkane-like peroxy chemistry. The following sections further sC4H9OH þ R % CH3CH2CH(CH2)OH þ RH (R43)
describe the important reactions in the oxidation of the butanol
isomers, including thermal decomposition, hydrogen abstractions, sC4H9OH þ R % CH3CHCH(CH3)OH þ RH (R44)
and low-temperature reaction pathways.
The unimolecular thermal decompositions of the butanol iso- sC4H9OH þ R % CH2CH2CH(CH3)OH þ RH (R45)
mers proceed primarily via dehydration reactions and CeC bond
scissions. CeO bond scissions are however negligible because their The H-atom abstraction rate coefficients by H radicals from the
BDEs are approximately 10 kcal higher than those of CeC bonds. butanols have not been measured experimentally. Ratkiewicz et al.
Rosado-Reyes and Tsang [387] used a single pulse-shock tube to [424] and Carstensen and Dean [302] presented ab initio calcula-
perform the first direct measurement of n-butanol dehydration. tions for all four isomers of butanol. The latter authors explained
The authors also estimated CeC bond scission rate coefficients that the OH group does not affect the rate coefficients of H-atom
using their measured GC-FID/MS results. They showed that previ- abstraction by H radical atoms after the b CeH site, and the rate rule
ous estimates used in combustion models [207,208,213] for these converges to that for hydrocarbons. Based on this assumption,
reactions were not in agreement with their rate coefficients. A Sarathy et al. [260] assigned the H-atom abstraction rate coefficient
notable conclusion from their work was the lack of pressure by H radicals from the a carbon and hydroxyl moiety based on the
dependence observed for these reaction rate coefficients (1.5e ethanol þ H work of Sivaramakrishnan et al. [366]. The b-carbon Ce
6.5 bar), which they verified by master equation calculations. Thus, H BDE in n-butanol is between that of a secondary and primary Ce
these reactions are at their high-pressure limit in practical com- H BDE of n-butane; the rate of abstraction was therefore calculated
bustion environments. Cai et al. [225] recently performed a theo- as the average of the two [212]. The abstraction rates for sites more
retical study on the unimolecular decomposition of n-butanol using than two carbons away from the hydroxyl group (e.g., g and d car-
variational transition state theory with master equation modeling bon atoms in n-butanol) were taken to be equal to analogous
to provide temperature- and pressure-dependent rate coefficients. abstraction rates from secondary and primary carbon sites of n-
At atmospheric pressure and in the temperature range of 850e butane [425].
1150 K, the rate coefficients for n-butanol dehydration calculated by Experimentalists and theoreticians have recently studied
Cai et al. [225] are a factor of 2e4 greater than those measured by hydrogen abstraction by OH radicals in all the butanol isomers. The
Rosado-Reyes and Tsang [387]. rate coefficients for all these isomers have been measured from low
The rates for thermal decomposition of iso-butanol [390], 2- to high temperatures, as shown in Table 5. The total rate of n-
butanol [388], and tert-butanol [203,204,386] have also been butanol þ OH at high temperatures (above 900 K) measured by
measured experimentally in single-pulse shock tubes. Rosado- Vasu et al. [380] and Pang et al. [381] are in good agreement with
Reyes et al. [390] showed that the dehydration rate coefficients each other. Subsequent high-temperature studies (above 900 K) by
of iso-butanol calculated by themselves and by Zhou et al. [422] Pang et al. on iso-butanol þ OH [382] and 2-butanol þ OH [383]
were within a factor of 2 of experimental measurements, and have also confined these total rate coefficients within narrow un-
that the pressure effects were minimal. The decomposition rate certainty limits. Atmospheric chemistry researchers have also
coefficient for 2-butanol was measured by Rosado-Reyes and studied butanol þ OH reactions at temperatures generally below
Tsang [388], and their measured dehydration rate coefficients 400 K. McGillen et al. [385] recently measured rate coefficients of
were in good agreement with the ab initio calculations of El-Nahas hydrogen abstraction by OH radicals for all four butanol isomers at
et al. [423]; however, the measured CeC bond scission rate co- low temperatures, and their results showed excellent agreement
efficients were notably different than those from calculations with previous measurements [373,376e379]. Thus, it has become
[423]. The measured rate coefficients for tert-butanol dehydration possible to estimate the total rate coefficient for all butanol isomers
by Tsang [203] were in excellent agreement with the experimental across low to high temperatures (220e1800 K), including the in-
measurements of Lewis et al. [204] and the rate coefficients termediate range of 400e1000 K where no measurements are
calculated from theory [386]. Rosado-Reyes et al. [390] also available. However, less certain are the branching ratios of
concluded that the rate coefficient of fuel alcohol dehydration hydrogen abstraction reactions from the a, b, g, and d CeH sites and
reactions is controlled primarily by the available number of the hydroxyl moiety, because direct measurements are not
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 83

available. Theoretical calculations [426e429] have been performed system were used to estimate abstraction rates for the other
to determine these branching ratios, but they are often unreliable butanol isomers. Their estimations derive abstraction rate co-
because their total rate coefficients do not match experimental data efficients based on the proximity of CeH sites to the hydroxyl
[260,385]. These reactions are particularly challenging to calculate moiety (e.g., a, b, g, etc.). In this way, the effects of hydrogen-
due to very low energy barriers, the large number of butanol con- bonded transition state complexes on the rate coefficients (e.g.,
formers, pre- and post-reaction complexes, and the theoretical entropy effects) are indirectly captured. The authors [260] found
treatment of the various hindered rotors. For this reason, Sarathy this approach to work well for iso- and 2- butanol, but not for tert-
et al. [260] indirectly estimated site-specific rate coefficients by butanol, wherein the rate expression for H-atom abstraction by
combining previous theoretical calculations, analogies with al- HO2 from a primary CeH bond of n-butane was used with the A-
kanes, and EvansePolyani correlations [411]. Their estimated rate factor modified to account for reaction path degeneracy. The rates
coefficients were shown to agree (i.e., within a factor of 2) with low- of hydrogen abstraction from the butanol isomers by HO2 were
and high-temperature rate coefficient measurements, and they consistent with those previously mentioned for methanol, ethanol,
enabled accurate simulations of low-temperature ignition delay and n-propanol, showing that abstraction from the a-site is
times sensitive to butanol þ OH reactions [260]. Recently, McGillen predominant.
et al. [385] (Fig. 36) employed an elegant approach of combining At high temperatures, the various C4H9O radicals formed from
experimental rate coefficient measurements, room temperature n-, 2-, iso-, and tert-butanol react via thermal decomposition (i.e.,
end-product yields, and modified structureeactivity relationships b-scission reactions) and unimolecular isomerization reactions.
(SAR) of Bethel et al. [430] to derive self-consistent, site-specific Isomerization reactions are favored for 5- and 6-member transition
rate coefficients for all four butanol isomers. The rates recom- state rings, which limits their importance to H-atom migration
mended by McGillen et al. [385] are yet to be tested in compre- reactions between the g or d site and the hydroxyl site (i.e., g-rad or
hensive combustion models of the butanol isomers. Future d-rad to O-rad). Isomerization reactions have been studied theo-
modeling studies should therefore make an effort to utilize them. retically in n-butanol’s C4H9O radicals by Zheng and Truhlar [431],
Hydrogen abstraction from butanol isomers by HO2 has never Xu et al. [432], and Zhang et al. [433], with reasonable agreement
been measured experimentally. Sarathy et al. [260] used a combi- between the studies. Experimental or theoretical studies of isom-
nation of theoretical calculations and estimations to determine erization reactions involving C4H9O radicals from 2-, iso-, and tert-
these rates. The H-atom abstraction rates by HO2 radicals were butanol were not found. Studies on b-scission reactions for various
based on the ab initio calculations (CCSD level of theory) for n- C4H9O radicals are also limited to the n-butanol system. Simmie
butanol þ HO2 by Zhou et al. [326]. Their calculations considered and Curran [434] theoretically studied the decomposition of bu-
primary effects (e.g., CeH BDEs) and secondary effects (e.g., pre- tanol’s a-hydroxybutyl decomposition to ethenol þ ethyl,
reaction complexes and transition states exhibiting hydrogen butanal þ H, or butanol þ H. Recently, Zhang et al. [433] performed
bonding) on the rate coefficient. Sarathy et al. [260] multiplied the a comprehensive theoretical study of the decomposition and
a-channel rate coefficient by a factor of 2.5, to obtain better isomerization of n-butanol’s hydrobutyl and butoxy radicals. Their
agreement with intermediate shock tube ignition delay experi- calculations showed that the channel leading to ethenol þ ethyl in
ments. EvansePolanyi type correlations with the n-butanol þ HO2 the a-hydroxybutyl radical is dominant at all temperatures, which

Fig. 36. Total and site-specific rate coefficients for butanol þ OH reactions proposed by McGillen et al. [385]. Black circles are low-temperature experimental measurements from
McGillen et al. [385] and red circles are high-temperature experimental data from Pang et al. [381e383]. Blue, green, and magenta circles represent end-product yields from various
studies, as described by Ref. [385].
Reprinted from McGillen et al. [385] with permission from ACS Publications.
84 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

agrees with the calculations of Simmie and Curran [434]. The b- followed by rapid decomposition. In the n-butanol system, the b-
radical decomposition was studied theoretically using addition of RO2 Waddington pathway leads to propanal, formaldehyde, and
OH to 1-butene as an entrance channel and subsequent decom- OH. These reactions are chain propagating and compete directly
position, either back to the reactants or on to methyl þ 2-propenol. with low-temperature chain branching pathways, thereby inhibit-
The calculated rate coefficient for OH þ 1-butene was in very good ing low-temperature reactivity. Sun et al. [344] conducted theo-
agreement with experimental studies at low temperatures [435]. retical calculations on these Waddington reactions in the tert-
Zhang et al. [433] found that the b- and g-hydroxybutyl radicals butanol and iso-butanol systems, showing their importance for the
decompose primarily to 1-butene þ OH and propene þ CH2OH, production of important carbonyl products.
respectively. Below 800 K, the d-radical can isomerize to form a The g- and d-hydroxybutyl radicals from n-, 2-, iso-, and tert-
butoxy radical, which then decomposes to propyl þ CH2O. At higher butanol also react with O2 to form stabilized RO2 radicals. These can
temperatures, the d-radical decomposes to ethylene þ CH2CH2OH. undergo concerted eliminations to form butenol þ HO2, cyclization
The pressure-dependent rate coefficients calculated by Zhang et al. to form epoxy alcohols, internal H-atom migration to form butyl-
[433] are recommended for use in chemical kinetic models for n- hydroperoxides (QOOH), and subsequent low-temperature chain
butanol. Similar comprehensive theoretical studies are needed for branching reactions. These pathways are largely unexplored, but
the other butanol isomers. can be included in models based on analogies with low-
At lower temperatures (below 1000 K), reactions between temperature oxidation schemes of alkanes. Welz et al. [346] also
C4H9O radicals (from n-, 2-, iso-, and tert-butanol) and O2 become identified a low-lying water elimination pathway for the con-
important. Due to the number of carbon atoms, the possible low- sumption of g- and d-QOOH radicals with a radical on the a site.
temperature reactions pathways in C4H9O þ O2 are numerous, Their importance in the low-temperature oxidation schemes for
including concerted elimination, stabilization of hydroxybutylper- alcohols has not been rigorously tested in butanol combustion
oxy (RO2) radicals, and subsequent isomerization, cyclization, and models. Detailed kinetic rate coefficient studies using state-of-the-
decomposition reactions. Welz et al. combined quantum-chemical art experimental and computational methods are needed to
calculations and experimental measurements of product formation develop accurate chemical kinetic models for low-temperature
to explore the low-temperature reaction mechanisms of hydroxy- oxidation of butanol and higher alcohols.
butyl radicals from n-butanol [346], iso-butanol [436], and tert- The butanol model utilized in the present analysis is that
butanol [436]. For example, Fig. 37 presents stationary points on developed by Sarathy et al. [260] for the four isomers of butanol
the PES for reactions of n-butanol’s a-hydroxybutyl and b- with minor modifications to unimolecular decomposition rate co-
hydroxybutyl radicals plus O2 with the most important low- efficients, as described recently by Vasu and Sarathy [238]. This
temperature products highlighted in red. The a-hydroxybutyl model has been comprehensively developed for combustion of
radical reacts with O2 to form butanal þ HO2, a reaction similar to butanol isomers at low and high temperatures. The model pre-
that discussed previously for a-hydroxyethyl radicals. This reaction sented in Refs. [232,233] has also been comprehensively tested
proceeds through a shallow well, implying that stabilization of the against iso-butanol data. In their paper [260], the authors present
a-hydroxybutylperoxy radical intermediate is only possible at high comparisons against shock tube ignition delay data, JSR speciation
pressures. This concerted elimination reaction leading to an data, premixed laminar flame speeds, and low-pressure premixed
aldehyde þ HO2 is the primary consumption pathway of a- laminar flame speciation data.
hydroxybutyl radicals derived from iso-butanol [436] and 2- The laminar flame speeds of n-, iso-, 2-, and tert-butanol/air
butanol. However, rate coefficients have not been measured mixtures have been measured in a number of studies, as noted in
experimentally or calculated. Table 6, and there is reasonably good agreement across the various
The b-hydroxybutyl radical from n-, 2-, iso-, and tert-butanol measurements obtained at atmospheric pressure. Recently, Wu
reacts with O2 to form the stabilized b-hydroxybutylperoxy radical and Law [236] performed a thorough experimental study that
(b-RO2), which is then subject to a variety of reactions. The most measured premixed laminar flame speeds of the four butanol
important of these is the Waddington reaction, which involves a isomers at 1, 2, and 5 bar. The authors utilized the chemical kinetic
six-membered ring isomerization to abstract an H atom from the model developed by Sarathy et al. [260] to interpret their results,
hydroxyl moiety (i.e., forming an alkylhydroperoxide alkoxy), as this model showed excellent agreement with the experimental

Fig. 37. PES for the a-hydroxybutyl þ O2 (left) and b-hydroxybutyl þ O2 (right) reactions. The most favorable pathways are highlighted in red (see main text in Ref. [346] for details).
Reprinted Welz et al. [346] with permission from ACS Publications.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 85

data (Fig. 38). The authors [236] found that n-butanol has the reactive, unsaturated branched intermediates in the case of iso-,
fastest flame speed, iso-butanol and 2-butanol have similar flame 2-, and tert-butanol compared to the less-reactive straight-chain
speeds, and tert-butanol has the slowest. Wu and Law [236] intermediates formed from n-butanol. The latter decomposes
further investigated the thermal and kinetic contributions to primarily to vinyl and methyl radicals, which accelerate flame
flame speeds to rationalize the observed differences. Comparable propagation. tert-Butanol and iso-butanol flames produce iso-
adiabatic flame temperatures and heat release profiles between butene and branched oxygenates (e.g., branched enols and
the fuels indicated that thermal effects cannot be responsible for acetone). iso-Butene is less reactive due to the formation of the
the variations in flame speeds. The underlying cause for differ- resonantly stabilized 2-methyl-allyl radical following H-atom
ences in laminar flame speeds is attributed to formation of less- abstraction, which results in lower flame speeds [109,417,437].
Similar branched unsaturated oxygenates formed in iso- and 2-
butanol flames results in lower concentrations of reactive vinyl
and methyl radicals and higher concentrations of less reactive allyl
radicals [109], thus lowering flame speeds.
The homogenous ignition delay of the butanol isomers at high
temperatures has been studied by a number of experimentalists, as
shown in Table 6. Recently, Stranic et al. [108] studied the ignition
of n-, iso-, 2-, and tert-butanol at high temperatures (i.e., above
1000 K) and pressures ranging from approximately 1e50 bar (4%
fuel in Ar/O2 mixtures). Both experimental data and simulation
results using the model of Sarathy et al. [260] showed that n-
butanol is the most reactive, followed by 2-butanol, with iso-
butanol exhibiting similar reactivity, while tert-butanol is much
less reactive. Molecular dehydration reactions are important in
controlling the relative reactivity of the butanol isomers under
these high-temperature, high-dilution conditions. tert-Butanol is
the least reactive because molecular dehydration reactions form
the stable iso-butene intermediate. The ignition of butanol isomers
at low to intermediate temperatures and high pressures has not
been studied extensively. Shock tube measurements of butanol/air
mixtures at temperatures below 1000 K and elevated pressures
have not been performed on the four isomers of butanol. n-Butanol
has been studied under such conditions by Heufer et al. [219],
Stranic et al. [223], and Zhu et al. [240].
Fig. 39 displays the measured and simulated n-butanol ignition
delay times using conventional filling methods [219,223] and the
constrained reaction volume (CRV) strategy developed by Hanson
et al. [101] and applied by Zhu et al. [240]. The data obtained
using conventional filling methods are simulated using variable
volumes profiles that account for experimental pre-ignition
pressure rise, while simulations of the CRV data are conducted
under constant pressure conditions. In both cases, the model is
able to simulate the temperature, pressure, and equivalence ratio
dependences. It is interesting to note that significant low-
temperature reactivity of n-butanol is observed at pressures
above 20 bar, a phenomenon not observed in the smaller alcohols.
Therefore, low-temperature peroxy chemistry is needed in n-
butanol chemical kinetic models to simulate engine-relevant
ignition behavior. The ignition of all four isomers of butanol at
low and intermediate temperatures and elevated pressures has
been studied in an RCM by Weber et al. [105], as shown in Fig. 40.
Under such conditions, n-butanol is notably more reactive than
iso-, 2-, and tert-butanol. Variable volume simulations using the
model by Sarathy et al. [260] are in good qualitative agreement
with the experimental data, but improvements in quantitative
agreement are needed. The study by Weber et al. [105] concluded
that low-temperature peroxy chemistry is important to all
butanol isomers, and accurate rate coefficients in this network are
needed. More details on low-temperature reaction pathways are
discussed later.
Species evolving from the oxidation and pyrolysis of the butanol
isomers have been studied in flow reactors, stirred reactors, non-
premixed flames, and premixed flames using various diagnostic
Fig. 38. Premixed laminar flame speeds for the four butanol isomers in air at 1, 2, and
5 bar and various of temperatures. Symbols are data taken by Wu and Law [236] and
techniques, as shown in Table 3. Most studies have focused on high-
lines are simulations. temperature oxidation in flames and pyrolysis reactors. Recent
Reprinted from Wu and Law [236] with permission from Elsevier. experimental and modeling studies by Cai et al. on n-butanol [225],
86 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Fig. 39. Measured ignition delay times in a ST (symbols) and simulations (lines). The left panel contains data from Refs. [219,223] at stoichiometric conditions using conventional
filling methods. The right panel contains data from Ref. [240] using the CRV method. The left panel is reprinted from Sarathy et al. [260] and the right panel is reprinted from Zhu
et al. [240] with permission from Elsevier.

2-butanol [235], and tert-butanol [224] have demonstrated the The high-temperature combustion of the butanol isomers differs
ability to reproduce high-temperature combustion and pyrolysis notably from the low-temperature consumption pathways; we
species. The authors combined quantum chemistry calculations to therefore present them separately. Fig. 41 presents the general high-
obtain rate coefficients for important unimolecular decomposition temperature reaction schemes for n-, iso, 2-, and tert-butanol. The
and b-scission reactions, and then compared their models with major product species arising from the direct decomposition of the
species profiles measured with vacuum ultraviolet photoionization fuels are in agreement with those measured in experiments, as
mass spectrometry (VUV-PI-MS). Sarathy et al. [260] and Frassol- shown in Table 7. At high temperatures, the principal consumption
dati et al. [318] have also developed combustion models of the four pathways are unimolecular decompositions (including de-
butanol isomers against VUV-PI-MS species profiles in low- hydrations) and hydrogen abstractions followed by b-scission of fuel
pressure premixed flames studied by Oßwald [221]. Low-pressure radicals. The primary stable decomposition products from n-butanol
premixed flames of n-butanol and iso-butanol have also been combustion are ethenol, 1-butene, 2-propenal, propene, ethylene,
probed by Hansen et al. [218,232] by VUV-PI-MS. The important along with the radicals ethyl, hydroxyl, methyl, hydroxymethyl, and
oxygenated hydrocarbon species arising from the combustion of a-hydroxyethyl. iso-Butanol yields the stable intermediates n-pro-
the four butanol isomers are shown in Table 7. All the butanol penol, n-propanal, iso-butene, and propene, plus the radicals
isomers produce formaldehyde, ketene, ethenol, and acetaldehyde. methyl, hydroxyl, and hydroxymethyl. 2-Butanol combustion is
The C3 and C4 species differ among the isomers due to the different notably more complex with stable products including iso-propenol,
molecular structures. The pathways leading to the different species acetone, ethenol, 1-butene, 2-butene, 1-propenol, n-propanal, and
are discussed next. ethylene, and the radicals methyl, ethyl, hydroxyl, and a-

Fig. 40. Measured ignition delay times in an RCM (symbols) [105] and simulations for the four isomers of butanol at 15 bar and 30 bar.
Reprinted from Weber and Sung [105] with permission from ACS Publications.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 87

Table 7
Important oxygenated hydrocarbon species measured [211,215,221,224,225,235,241] and predicted [121,218,224,225,232,233,235,260,318] from the high-temperature com-
bustion of the four butanol isomers.

n-Butanol 2-Butanol iso-Butanol tert-Butanol

C1 Formaldehyde Formaldehyde Formaldehyde Formaldehyde


C2 Ketene, ethanol, acetaldehyde Ketene, ethanol, acetaldehyde Ketene, ethanol, acetaldehyde Ketene, ethanol, acetaldehyde
C3 2-Propenol 1- Propenol, n-propanal, acetone, iso-propenol n-Propenol, propanal, 2-propenol iso-Propenol, acetone
C4 n-Butanal, butenol isomers 2-Butanone, butenol isomers 2-Methylpropanal, butenol isomers e

hydroxyethyl. tert-Butanol consumption is relatively simple due to their high BDEs. In addition, the b-hydroxybutyl radical favors the
the high degree of symmetry. iso-Propenol, acetone, and iso-butene OH propagation (Waddington mechanism) route because internal
are the important stable products, plus methyl radicals. H-atom migration reaction leading to QOOH (and subsequent chain
The low-temperature oxidation of the butanol isomers is shown branching) can only proceed via high-energy barrier 5-membered
in Fig. 42 to help explain the differences in low-temperature ring transition states. The low-temperature pathways for 2-
reactivity discussed previously. Following initial H-atom abstrac- butanol show the same chain termination pathways for reactions
tion by OH, the subsequent possible low-temperature reaction involving its a-hydroxybutyl radical. The two b sites (i.e., one of the
pathways for the butanol isomers distinguish their reactivity. n- ethyl branch and one on the methyl branch) contribute little to OH
Butanol radicals react with molecular oxygen and then can either branching because of competing OH termination and OH propa-
lead to OH branching (via QOOH), OH propagation (via the Wad- gation routes. OH branching comes mainly from the g-hydroxy-
dington mechanism), or OH termination (aldehyde/enol þ HO2). It butyl radical. tert-Butanol is unique because there are no a-CeH
should be noted that some of the QOOH is used to form one OH bonds. The three primary b-radicals can react with O2 and lead to
radical and one HO2 radical that is also used to form low- OH branching and/or OH propagation. OH propagation via the
temperature radical branching. The g- and d-hydroxybutyl radi- Waddington mechanism is favored because the high BDEs of the
cals contribute most to OH branching in n-butanol. The a-hydroxy- primary CeH bonds slow internal H-atom migration reactions.
butyl radical leads primarily to OH termination, and the b-
hydroxybutyl radical leads to either OH branching, OH termination, 4.9. The combustion chemistry of higher alcohols
or OH propagation. iso-Butanol is less reactive than n-butanol
because chain branching evolves from g-hydroxybutyl radicals; The combustion chemistry of higher alcohols (i.e., C5 and larger)
initial H-atom abstractions are slower from g primary sites due to has not been studied extensively. Several recent works in the past

Fig. 41. Reaction path diagrams for high-temperature combustion of (a) n-butanol, (b) iso-butanol, (c) tert-butanol, and (d) 2-butanol. R denotes a radical, such as OH, HO2, and H.
88 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Fig. 42. Simplified schematics of important low-temperature reaction pathways for the four isomers of butanol: (a) n-butanol, (b) iso-butanol, (c) 2-butanol, and (d) tert-butanol.
Reprinted from Sarathy et al. [260] with permission from Elsevier.

3e4 years have focused on the isomers of pentanol and n-hexanol. other alcohols, dehydration reactions leading the formation
Togbé et al. conducted experiments on n-pentanol [246] and n- of water and an alkene are important from 900 K and above, whilst
hexanol [244] in JSRs and premixed laminar flames, and then CeC and CeO bond scissions are competitive at higher tempera-
developed chemical kinetic models to simulate their experimental tures (above 1200 K). Given the rather limited fundamental kinetic
data. n-Pentanol oxidation in the JSR did not exhibit any cool flame data available, Heufer et al. [248] and Sarathy et al. [252] developed
behavior; low-temperature kinetic pathways were therefore not comprehensive chemical kinetic models based on analogies to al-
included in their model. On the other hand, n-hexanol has cool kanes and alcohols of similar structures. The development of re-
flame reactivity and a low-temperature oxidation scheme was action classes follows the methods described in Section 4.3 and
assembled to reproduce this behavior in the JSR. Neither of the Box 1. This approach was first verified by quantum chemical cal-
models by Togbé et al. [244,246] have been compared against to culations for CeH, CeC, and CeO bond dissociation energies that
low-temperature ignition data. Recently, Heufer et al. [248] and showed higher alcohols can be sub-divided into an alcohol-specific
Sarathy et al. [252] developed comprehensive models for n-pen- and alkane-like portion, as shown in Fig. 43 for n-pentanol. All
tanol and iso-pentanol (3-methyl-1-butanol), respectively, which reaction rate coefficients were assigned based on analogous pri-
were compared against ignition delay measurements in RCMs and mary and secondary alkane (e.g., n-butane and iso-butane) sites and
shock tubes, speciation profiles in JSRs, and premixed laminar a, b, and g alcohol sites (e.g., n-butanol and iso-butanol).
flame speeds. These pentanol isomer models form the basis for the The low-temperature reaction mechanism for pentanol isomers
present discussion on the combustion chemistry of higher alcohols. is also assembled based on direct analogy with butanol isomers.
An extensive discussion on fundamental kinetic studies of high- Recent experimental and theoretical studies by Welz et al.
temperature reactions (e.g., hydrogen abstraction, b-scission, [345,439] enabled the addition of several new low-temperature
radical isomerization) and low-temperature reaction pathways is reaction pathways to the models. Welz et al. [345] described
not possible for the pentanol isomers because such studies are Waddington-type pathways involving hydrogen transfer from the
absent from the literature. Zhao et al. [438] employed quantum OH group as important for the g-ROO and d-ROO radicals via 7- and
chemical calculations with master equation simulations to derive 8-membered transition state rings. These reaction pathways were
pressure-dependent rate coefficients for the unimolecular decom- therefore added to the models using estimated rate coefficients
position of three pentanol isomers (n-pentanol, 3-methyl-1- [252]. Welz et al. [439] also presented a water elimination pathway
butanol, and 2-methyl-1-butanol). As shown previously for the from hydroxypentylhydroperoxide (QOOH) radicals where the
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 89

Fig. 43. Premixed laminar flame speeds for n-pentanol/air (left) and iso-pentanol/air (right) at 1 bar. (Left) n-Pentanol symbols are 423 K data presented in Togbé et al. [246].
Reprinted from Heufer et al. [248] with permission from ACS Publications. (Right) iso-Pentanol symbols are 353 K data presented in Sarathy et al. [252] and iso-butanol symbols are
353 K data presented in Liu et al. [228]. Reprinted from Sarathy et al. [252] with permission from Elsevier.

hydroperoxide is on the g-site and the radical is on the a-site. This ignition delay, are similar. iso-Pentanol measurements and simu-
reaction pathway is also adopted in the iso-pentanol model lations at a temperature of 353 K are also presented in Fig. 44.
developed by Sarathy et al. using estimated rate coefficients [252]. Similar to the case of normal alcohols, the data and model indicate
The laminar flame speeds of n- and iso-pentanol/air mixtures that iso-pentanol and iso-butanol have similar flame speeds. Again
have been measured in few studies, as noted in Table 3. n-Pentanol this is due to the fact that the intermediate pools of radical species
premixed laminar flames speeds were measured by Togbé et al. are similar in iso-pentanol and iso-butanol flames.
[246] at 423 K, and simulations using the present chemical kinetic The homogenous ignition delay of the pentanol isomers has
model are shown in Fig. 44. The model indicates that the laminar been studied in STs and RCMs, as shown in Table 3. Recently, Heufer
flame speed is very similar for n-pentanol and n-butanol over the et al. [103,108] studied the ignition of n-pentanol at temperature
whole range of equivalence ratios, a phenomenon which has also from 640 to 1200 K and pressures from 9 to 30 bar. The 30 bar ST
been observed for C5eC12 n-alkanes [440]. The concentrations of ignition delay data were directly comparable to data obtained for n-
reactive H, OH, and CH3 radicals generated during the combustion butanol under similar conditions [219], as shown in Fig. 45. Both the
of n-butanol and n-pentanol are similar, and thus their global high experiments and present chemical kinetic model indicate that n-
temperature reactivities, manifested as premixed flame speed or pentanol has greater reactivity in the NTC region, which is

Fig. 44. Ignition delay times in an ST and RCM for n-pentanol (left) and iso-pentanol (right). (Left) ST data for n-butanol from Ref. [219] and n-pentanol [103] at 30 bar, stoi-
chiometric conditions. Reprinted from Heufer et al. [248] with permission from ACS Publications. (Right) Data for iso-pentanol from Sarathy et al. [252] at 20 bar and range of
equivalence ratios. Reprinted from Sarathy et al. [252] with permission from Elsevier.
90 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Fig. 45. Shock tube reaction path analysis for n-butanol and n-pentanol at 700 K, 30 bar, and 4 ¼ 1.0. The reaction fluxes are given for 20% fuel consumption.
Reprinted from Heufer et al. [248] with permission from ACS Publications.

attributed to its longer chain length promoting low-temperature pentanol isomers under specific low-temperature conditions. A
chain branching reactions. Above 1000 K, n-pentanol and n- reaction path analysis for n-butanol and n-pentanol under similar
butanol exhibit similar high temperature reactivity, which is shock tube ignition delay conditions is presented in Fig. 46 to
consistent with the flame speed comparison presented above. iso- highlight the differences in their reactivity. It was previously shown
Pentanol ignition delay measurements are available at pressures that the n-pentanol exhibits greater low-temperature reactivity
ranging from 7 to 60 bar [249,252], and model simulations near than n-butanol in shock tube ignition delay experiments. Both fuels
20 bar are presented in Fig. 45. The data spanning lean to rich undergo typical H-atom abstraction reactions followed by O2
conditions show a strong equivalence ratio dependence at low addition to form RO2 radicals. The primary driver for greater reac-
temperatures that is less prominent at higher temperatures. The tivity in n-pentanol is its longer carbon chain length, which permits
present chemical kinetic model simulations are able to capture the hydrogen atom abstraction on a secondary “alkane-like” carbon.
low-, NTC-, and high-temperature reactivity of iso-pentanol in the The subsequent d-ROO undergoes a six-membered ring internal
ST and RCM. isomerization to form QOOH, which later reacts with O2 and leads
Speciation measurements resulting from the oxidation of n- and to significant chain branching. Other minor pathways contributing
iso-pentanol are available in JSRs. Researchers at CNRS Orleans have to the differences in low-temperature reactivity between n-butanol
presented data at a range from lean to rich equivalence ratios for n- and n-pentanol are discussed by Heufer et al. [248]. For iso-pen-
pentanol near 10 bar [246] and iso-pentanol near 10 bar [245] and tanol, an NTC region reaction path analysis under JSR and ST con-
5 bar [252]. The chemical kinetic models for each fuel are able to ditions is presented. In general, the reaction pathways are similar to
qualitatively reproduce the transient fuel reactivity, formation of those presented and discussed previously for n- and iso-butanol.
major product species, and profiles of oxygenated intermediates. In Similar to the other alcohols, a-hydroxypentyl radicals react with
general, the important oxygenated products are identical to those O2 to generate an aldehyde (i.e., iso-pentanal) and HO2. The other
presented previously for n-butanol and iso-butanol (Table 7), albeit hydroxypentyl radicals react with O2 to form hydroxyalkyl peroxy
C5 aldehydes are also produced from the reaction of a-hydroxy- (ROO) radicals. These radicals are mainly isomerized to hydrox-
pentyl radicals with molecular oxygen. The chemical kinetic model yalkyl hydroperoxide (QOOH) or undergo a 5-centered concerted
for iso-pentanol [252] indicates that the acetone present at low and reaction to form HO2 and an enol species. Waddington type re-
intermediate temperatures arises from the water elimination actions are shown to be important for b-ROO radicals. The con-
pathway proposed by Welz et al. [439]. Both experiments and sumption of a-hydroxypentyl hydroperoxide radicals results in the
models indicate that the pentanol isomers do not display low- and formation of a cyclic ether plus OH (minor route in STs) or a
NTC-temperature reactivity in the JSR under the conditions studied concerted elimination to carbonyl hydroperoxide plus HO2 (major
(1000 ppm fuel in air), which is a behavior consistent with smaller route in STs). The water elimination pathway is shown to play an
alcohols; however, Togbé et al. [244] showed that low- and NTC- important role during JSR oxidation, but it is not important during
temperature reactivity is observed for n-hexanol. ignition delay under ST conditions.
Extensive studies on the reaction pathways for combustion of
pentanol isomers are lacking, so presenting general high- and low- 5. Insights into engine-relevant phenomena
temperature mechanistic pathways is not warranted. The high
temperature decomposition pathways are analogous to those A primary goal of this article is to use the knowledge gained on
already presented for the butanol isomers. For simplicity, we pre- fundamental combustion chemistry of alcohols toward under-
sent actual pathway analyses performed for the combustion of standing their performance in internal combustion engines. In this
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 91

Fig. 46. Reaction path analysis for iso-pentanol oxidation in a JSR and ST. Black JSR 5 atm/850 K/italic (4 ¼ 0.35) bold(4 ¼ 1.0) plain (4 ¼ 4.0). Gray ST 20 atm/800 K/4 ¼ 1 at 20% fuel
consumed.
Reprinted from Sarathy et al. [252] with permission from Elsevier.

section, the comprehensive C1eC5 alcohol kinetic model is utilized which are comparable to those of n-alkanes (ethane, propane, n-
to predict the effects of alcohol fuel structure on laminar flame butane, n-pentane, and n-heptane). All the alcohols and alkanes
speed, homogenous ignition delay time, and combustion products. have similar stoichiometric fuel/air adiabiatic flame temperature,
These are then related to engine-relevant phenomena, specifically so differences in laminar flame speed are attributed to fuel com-
pre-ignition and super-knock, fuel autoignition, and pollutant bustion chemistry. Chemical kinetic modeling reveals that meth-
emissions. anol is the most reactive in premixed flames because high-
Kalghatgi [88,441] explains that pre-ignition is an abnormal temperature decomposition reactions (i.e., hydrogen abstraction
combustion phenomenon that occurs in turbocharged DI SI en- and b-scission) primarily produce formaldehyde, which de-
gines, wherein a propagating flame front is established prior to the composes quickly to reactive H radicals. Normal C2eC5 alcohols
spark plug firing. Pre-ignition is different than traditional knock, decompose to intermediates that populate the H-radical pool at
which is an autoignition phenomenon caused by temperature and approximately the same rate, so these fuels display similar laminar
pressure increasing in the end-gas ahead of the spark-initiated flame speeds. It can be concluded with a high degree of confidence
flame front. Pre-ignition may cause the temperature and pressure that larger normal alcohols (i.e., >C6) will display similar laminar
of the end-gas to increase faster than it would under conventional flame speeds. Substitution of methyl or OH in C3eC5 alcohols (i.e.,
spark timing, and thus may lead to a heavy knock or “super-knock” branching) leads to a reduction in flame speed when compared to
event. The criteria for pre-ignition to occur are a local hot spot that normal alcohols, as is also shown for iso-alkanes. For example, tert-
initiates thermal runaway followed by a propagating flame front. butanol/air and iso-octane/air mixtures display dramatically sup-
The probability of pre-ignition increases as as the fuel/air mixture’s pressed laminar flame speed due to multiple substitutions.
laminar flame speed increases [88,441]. Substituted molecules are not as effective at populating the H
Fig. 47 displays simulated laminar flame speeds for alcohol/air radical pool because their decomposition leads to the formation of
mixtures using the C1eC5 alcohol kinetic model presented herein. resonantly stabilized radical intermediates. A complete under-
The figure also presents simulated laminar flame speeds for stoi- standing on the effects of substitution (i.e., location and degree) on
chiometric fuel/air mixtures of C1eC5 alkanes, n-heptane, and iso- combustion of alcohols is not yet available due to the lack of
octane using the kinetic model of Mehl et al. [86,442]. Amongst the experimental data and chemical kinetic models for the many iso-
alcohols, methanol has the highest laminar flame speed, followed mers of C5 and greater alcohol fuels.
by ethanol. The remaining C3eC5 normal alcohols (i.e., n-propanol, The aforementioned analysis suggests that methanol and
n-butanol, and n-pentanol) display similar laminar flame speeds, ethanol will increase pre-ignition propensity due to their higher
92 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Fig. 47. Simulated laminar flame speeds for C1eC5 alcohols in air at 1 atm, 353 K (left) and laminar flame speeds for various alcohols and hydrocarbons in air at 1 atm, 353 K, 4 ¼ 1.1.

laminar flame speeds. Higher molecular weight normal and ignition delay time at 825 K and 25 atm. The ignition delay time of
branched alcohols will not increase pre-ignition propensity PRF fuels are simulated using the comprehensively tested gasoline
compared to typical n- and iso-alkanes found in conventional surrogates model by Mehl et al. [86,442]. n-Heptane has a RON of
gasoline fuels. Although higher laminar flame speeds can increase 0 and iso-octane has a RON of 100, and a non-linear relationship is
pre-ignition propensity, other fuel factors need to be considered. observed versus the predicted homogenous ignition delay time.
For example, low-molecular-weight alcohols have higher latent Methanol, ethanol, the propanol isomers, and tert-butanol have
heats of vaporization, which can reduce in-cylinder temperatures RON values greater than 100 and predicted ignition delay times
and decrease pre-ignition propensity. Furthermore, an increase in longer than those of PRF100. However, n-propanol is predicted as
pre-ignition propensity does not necessarily result in a super-knock less reactive than ethanol despite its lower RON (104 versus 109,
event. As will be shown next, higher octane fuels with longer ho- respectively). n-, iso-, and 2-Butanol isomers have octane ratings
mogenous ignition delay times decrease knock propensity. Thus, similar to that of conventional gasoline (90e105) and shorter
although methanol and ethanol exhibit higher laminar flame ignition delay times than PRF100. n-Pentanol has a RON of 82 and
speeds that increase pre-ignition propensity, their high octane the model predicts its decreased reactivity compared to the other
numbers can inhibit the occurrence of super-knock. alcohols. The measured RON of iso-pentanol (3-methyl-1-butanol)
The autoignition of fuel/air mixtures is an important design is 93, but the model predicts an ignition delay time closer to
consideration in SI, CI, and HCCI engines. Premixed fuel/air auto-
ignition in SI engines, referred to as knock, occurs in the end-gas
region due to increasing in-cylinder temperature and pressure
following spark plug firing. A fuel’s resistance to autoignition under
SI engine conditions is quantified by its RON and MON values.
Higher RON and MON values indicate lower autoignition pro-
pensity. In CI and LTC engines fuel/air autoignition is desired for
controlling combustion timing and susbequent heat release. The
cetane number of a fuel quantifies its autoignition propensity under
CI engine conditions; higher cetane numbers indicate higher pro-
pensity for autoignition. There is an inverse correlation between CN
and octane number (RON or MON). Fuels with RON < w60 and
CN > w30 are suitable for CI engines [88,443], those with
RON > w60e80 are suitable for LTC engines [444e447], and fuels
with RON > w80 are appropriate for SI engines. It was previously
stated that premixed fuel/air ignition delay time can be correlated
with its octane rating [86,87].
ST and RCM ignition delay time measurements at high pressures
(e.g., greater than 10 bar) and low to NTC temperatures (e.g., 600e
1000 K) correlate better with fuel octane and cetane ratings than
measurements at lower pressures and/or higher temperatures.
Sarathy et al. [87] have proposed that the RON of normal and
branched octane isomers can be closely correlated with their ST
ignition delay time under specific conditions (e.g., 825 K and
25 atm). Fig. 48 presents the measured RONs of C1 to C5 alcohol Fig. 48. Correlation of research octane number (RON) with ignition delay time for
fuels (from Table 1), as well as the primary reference fuel (PRF) stoichiometric fuel/air mixtures near 835 K and 20 atm. The line represents PRF
mixtures of n-heptane and iso-octane, versus simulated fuel/air mixtures of n-heptane and iso-octane.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 93

PRF100. The aforementioned discrepancies for n-propanol and iso- related works in the literature, enable a lucid understanding of
pentanol suggest that additional RON measurements are needed to how alcohol fuels reduce soot and particulate matter during en-
reduce the uncertainties in applying the correlation, and more gine combustion. Fig. 49 from McEnally and Pfefferle [448] pre-
fundamental ST or RCM data is required to further refine detailed sents the measured yield sooting idex (YSI) for various alcohols,
chemical kinetic models. The high sensitivity of alcohols fuels n-alkanes, and ethers as a function of their carbon number. The
causes the correlation for alcohols to be quantitatively different alcohol data is presented for linear alcohols with various OH
than that for PRF fuels, which have zero sensitivity by definition. substitutions (e.g., n-butanol is a 1-alcohol, 2-butanol is a 2-
Higher molecular weight normal alcohols (C5 and higher) have alcohol and 3-pentanol is a 3-alcohol). The YSI is determined
lower octane ratings and display significant low- and NTC- by doping an atmospheric pressure non-premixed methane/air
temperature reactivity, thus making them more suitable for CI coflow diffusion flame with 1000 ppm of the test compound and
engine applications. C1eC4 alcohols have higher octane numbers measuring the maximum soot mass concentration with laser-
and are more applicable to SI engines. The high sensitivity of al- induced incandescence. The measured soot mass concentrations
cohols fuels will improve performance in advanced engines, such as are normalized to a YSI using n-hexane as YSI ¼ 0 and benzene as
LTC and turbocharge DI SI engines [88,443]. An alcohol’s ignition YSI ¼ 100 (i.e., analogous to an octane rating) [448]. The domi-
propensity is closely related to its molecular structure. Hydrogen nant fuel consumption pathways in the YSI flames are high-
abstraction from alcohol fuels is favored at the a-site (i.e., the C temperature thermal decomposition reactions. Westbrook
bonded to the OH group) due to its proximity to the OH moiety and et al.’s chemical kinetic modeling study of soot formation in
lower BDE. The subsequent a-hydroxyalkyl þ O2 reaction in alcohol diesel engines has shown that soot is formed in rich combustion
fuels suppresses low-temperature reactivity by competing with regions [83], wherein fuel consumption is dominated by high-
conventional low-temperature radical chain propagating and temperature thermal decomposition reactions. Thus, the YSI
branching reactions that initiate ignition processes. This inhibition measurements are useful for understanding fuel structure effects
of low- and NTC-temperature reactivity is the cause of high octane on soot emissions from engines.
sensitivity in alcohol fuels. The branching ratio of forming the a- The results in Fig. 49 [448] indicate that the YSI of linear alcohols
hydroxyalkyl radical decreases as the carbon chain length in- increases as the carbon number increases, which is also true for n-
creases, so its inhibiting effects on ignition are lesser for larger al- alkanes. Higher carbon number fuels increase soot formation
cohols. It is postulated that the critical chain length at which the OH because they inevitably form larger alkenes via unimolecular
moiety no longer affects low- and NTC-temperature ignition delay decomposition (i.e., Reaction class 1) and radical b-scission re-
is 7e8 carbon atoms. Therefore, n-heptanol or n-octanol are ex- actions (i.e., Reaction class 4). These large alkenes subsequently
pected to have ignition delay times and octane numbers similar to participate in PAH growth and soot formation processes [448,449].
n-heptane and n-octane, respectively, but further research is The present chemical kinetic model has demonstrated that alcohol
needed on higher molecular weight alcohols to confirm this. combustion at high temperatures directly results in the formation
Methyl and OH substitution in alcohols tends to increase ignition of large alkenes via four-center unimolecular decomposition re-
delay times (i.e., increase the octane number) because substitutions actions [206] (i.e., Reaction class 1 and Fig. 15). These reactions also
increase the number of primary CeH sites, which have higher BDEs sequester the fuel bound oxygen atom into a water molecule,
and suppress low-temperature radical chain propagating and thereby preventing the oxidation of carbon atoms and allowing
branching reactions. them to participate in soot growth processes. Therefore, C4 and
Fundamental speciation studies presented in this manuscript larger alcohol fuels have higher sooting tendencies compared to n-
confirmed that carbonyl emissions commonly observed in engine alkanes of the same carbon number. OH substitution in alcohols
research are due to the hydroxyl moiety in alcohol fuels. The increases sooting tendency because the faster rates of four-center
important oxygenated emissions from C1eC5 alcohols are form- unimolecular decomposition reactions yield more large alkenes.
aldehyde, acetaldehyde (i.e., ethanal), propanal acetone, butanal McEnally and Pfefferle [448] explain that alcohol fuels exhibit
isomers, and pentanal isomers. The production of these species is lower soot emissions in engines studies because they are used to
directly linked with molecular structure and combustion chemis-
try pathways. At high-temperature stoichiometric or fuel-lean
conditions, carbonyl products are formed primarily via hydrogen
abstraction from the a-site and subsequent b-scission reactions.
Methanol’s a-hydroxymethyl radical produces formaldehyde,
whereas ethanol’s a-hydroxymethyl produces acetaldehyde. The
a-hydroxyalkyl radicals in C3eC5 normal alcohols form ethenol,
which quickly isomerizes/tautomerizes to form acetaldehyde. The
high temperature formation of carbonyl compounds may be
averted at fuel-rich conditions, wherein hydrogen abstraction re-
actions are less favorable and unimolecular water elimination re-
actions favor the formation of alkenes. The C1eC5 alcohols also
produce carbonyls directly via a-hydroxyalkyl þ O2 % aldehyde/
ketone þ HO2 reaction. At low and intermediate temperatures, this
is the primary pathway to production of aldehydes and ketones,
especially the higher molecular weight carbonyls found during
combustion of C3 and larger alcohols.
As noted in Section 1, fueling engines with alcohols decreases
soot and particulate matter compared to petroleum hydrocarbon
fuels. The chemical kinetic modeling work presented in this
Fig. 49. Yield sooting index (YSI) of alcohols, n-alkanes, and ether as a function of
article does not enable the direct simulation of soot emissions carbon number. Symbols are measured data points and the line is a quadratic least-
during alcohol fuel combustion. Nevertheless, the present squares fit data for n-alkanes.
model’s insights into fuel decomposition pathways, together with Reprinted from McEnally and Pfefferle [448] with permission from ACS Publications.
94 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

Box 2
Future directions for alcohol combustion chemistry research NOx (i.e., prompt and thermal) under idealized condi-
tions, such that comprehensive chemical kinetic models
can be developed with PAH growth [460,461] and NOx
- Detailed insights into alcohol combustion chemistry are formation [462] sub-mechanisms. These models can then
primarily driven by experimental observations, which be used to simulate soot in more complex reacting flows,
enable the development of comprehensive chemical ki- such as laminar non-premixed flames [463e465], turbu-
netic models. Additional fundamental experimental lent non-premixed flames [466,467], and eventually
measurements are needed to better understand alcohol practical engine systems.
combustion under conditions of relevance to internal - Additional fundamental studies are needed to determine
combustion engines. These include laminar flame speed the effects of alcohol blending with conventional petro-
measurements at higher pressures (5e30 bar) and un- leum hydrocarbon fuels. This paper highlighted several
burnt gas temperatures (400e600 K) and ignition delay fundamental studies on alcohol/PRF blends, but further
time measurements at higher pressures (10e100 bar) and studies with multicomponent surrogates and real trans-
lower temperatures (550e950 K). Speciation measure- portation fuels are required to better understand the
ments in flames and flow reactors have been largely synergistic and/or antagonistic effects of alcohol
restricted to high temperatures and low pressures. Ef- blending on important combustion properties (e.g., oc-
forts are needed to better capture alcohol combustion tane number [468]).
intermediates at elevated pressures (>1 bar) and low- to - Chemical kinetic modeling and fundamental experiments
intermediate-temperatures (550e950 K). should be utilized to enable a deeper understanding of
- Advances are needed in chemical kinetic model devel- alcohol combustion chemistry in engines. Important
opment to reduce uncertainties in the large number of phenomena of relevance to future alcohol-fueled engines
reactions and rate constant parameters. For example, (e.g., turbocharged DI SI and lean-burn LTC) and
chemical kinetic models for iso-butanol combustion requiring more fundamental research are the effects of
[233,260] with widely discrepant reaction rate constants exhaust gas recirculation and negative valve overlap [469
are able to accurately simulate historically measured e472], thermal and mixture stratification [473e475], in-
data. This indicates a need to advance experimental and termediate temperature heat release [90,97,476], the
theoretical methods to further constrain kinetic model chemical origins of octane sensitivity [477,478], and the
parameters. pressure [443] and equivalence ratio sensitivity [474,475]
- Theoretical studies exploring the potential energy sur- of autoignition.
face of low-temperature alcohol oxidation can reveal new
reaction classes (e.g. [345,346,436,439]). As shown in this
paper, theoretical estimates of reaction rate constants displace higher carbon number n-alkane fuels, which invariably
can be readily adopted to improve chemical kinetic have higher sooting tendencies. The fuel bound oxygen atom does
models, but improvements in even the highest level not actually have any chemical benefit toward reducing soot; it’s
theoretical calculations presently available are needed primary role is the modification of physical properties that enables
[410,451]. Theoretical tools should be applied to calculate smaller hydrocarbon chains and higher H/C ratio structures to be
rates for low-temperature reaction pathways and site- blended as liquid fuels [448]. Displacing high carbon number n-
specific hydrogen atom abstraction reactions. alkanes (e.g., n-heptane) and highly sooting aromatics [450] with a
- Undoubtedly, there is a need for elementary kinetic rate small carbon number alcohols (e.g., ethanol) decreases soot emis-
measurments for alcohol specific reactions. These sions from internal combustion engines.
experimental studies have seen a decline in the recent
years, but resources must be allocated to continue such 6. Summary and conclusions
efforts and advance them [369,384]. Measurements at
combustion relevant temperatures (600e1400 K) are The flammability of alcohols has been known for many cen-
needed to ensure the accuracy of reaction rate constants turies, and their combustion for energy dates back more than a
used in chemical kinetic models. hundred years. Despite this long history, a lucid understanding of
- In addition to the aforementioned experimental and alcohol combustion chemistry has only become available in the
theoretical studies, alcohol combustion modeling can be three decades preceding this article’s publication date. Literature
improved using uncertainty quantification/optimization since the early 1930s was surveyed to provide a comprehensive
tools [452e459]. These approaches can develop predic- summary of experimental and theoretical studies. A rapid growth
tive models needed to advance the design of energy in alcohol combustion chemistry in recent years can be attributed
systems. to the improvement of experimental facilities, development of
- The determination of premixed fuel/air laminar flame innovative diagnostic techniques, and advances in computational
speeds is continuously evolving with uncertainties being hardware and software.
reduced. Improved experimental methods and the anal- This article’s compiled knowledge of alcohol combustion
ysis and/or interpretation of measured data are needed to chemistry enabled a better understanding of combustion in en-
provide accurate laminar flame speed values for alcohols gines. Engine-relevant combustion phenomena considered were
at a range of pressures, temperatures, and equivalence the fuel’s ignition delay time, laminar flame speed, and emissions
ratios. characteristics. These combustion features were analyzed by
- An important area of practical relevance largely missing surveying fundamental combustion experiments in premixed
in the scientific literature is experimental and modeling laminar flames, batch reactors (i.e., STs and RCMs), and flow re-
studies on NOx and soot formation during alcohol com- actors (i.e., JSRs). This manuscript highlighted the progress made
bustion. Basic research should attempt to measure the toward understanding detailed chemistry during low- and high-
effects of alcohols on formation of soot precursors and temperature oxidation; with new experimental targets and new
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 95

theoretical insights, it is now possible to assemble semi-predictive [6] Hansen AC, Zhang Q, Lyne PWL. Ethanol-diesel fuel blends e a review. Biores
Technol 2005;96:277e85.
models for alcohols fuels. Thus, a consistent and comprehensive
[7] Agarwal AK. Biofuels (alcohols and biodiesel) applications as fuels for in-
chemical kinetic model for C1eC5 alcohols was compared with the ternal combustion engines. Prog Energy Combust Sci 2007;33:233e71.
experimental data, and then used to understand chemical con- [8] International Energy Agency. World energy outlook 2012 e renewable en-
tributions toward the aforementioned global combustion ergy outlook; 2013.
[9] International Energy Agency. World energy outlook factsheet; 2012.
properties. [10] World fuel ethanol production. Renewable Fuels Association; http://
The detailed understanding of alcohol combustion chemistry ethanolrfa.org.
presented in this article provides a fundamental explanation of [11] Levey M. Babylonian chemistry: a study of Arabic and second millenium B.C.
perfumery. The University of Chicago Press; 1956.
why these are fuels suitable for internal combustion engines. [12] Dietler M. Alcohol: anthropological/archaeological perspectives. Annu Rev
Higher molecular weight alcohols (e.g., C4 and larger) are benefi- Anthropol 2006;35:229e49.
cial for practical applications due to their higher energy densities. [13] Allchin FR. India: the ancient home of distillation? Man 1979;14:55e63.
[14] Hassan AY, Ahmed M, Iskandar AZ. Science and technology in Islam. UNESCO
However, ignition delay times and octane numbers are lower for Publications; 2001.
larger alcohols, making them more suitable for CI engines. Methyl [15] Forbes RJ. Short history of the art of distillation, from the beginnings up to
or OH substituted alcohols of the same carbon number have the death of cellier blumenthal. E. J. Brill; 1948.
[16] Peralta-Yahya P, Keasling J. Advanced biofuel production in microbes. Bio-
higher octane values, so are inherently better SI engine fuels. technol J 2010;5:147e62.
Advanced LTC engine concepts may utilize the unique chemical [17] Shen CR, Liao JC. Metabolic engineering of Escherichia coli for 1-butanol and 1-
features of alcohols to control ignition processes, improve com- propanol production via the keto-acid pathways. Metab Eng 2008;10:312e20.
[18] Jones DT, Woods DR. Acetone-butanol fermentation revisited. Microbiol Rev
bustion performance, and reduce emissions. From a fuel produc-
1986;50:484e524.
tion perspective, engineers should strive toward producing [19] Nigam PS, Singh A. Production of liquid biofuels from renewable resources.
alcohols spanning a range of carbon numbers for use in SI, CI, and Prog Energy Combust Sci 2011;37:52e68.
LTC engines. Conventional refining processes (e.g., catalytic [20] Lee SY, Park JH, Jang SH, Nielsen LK, Kim J, Jung KS. Fermentative butanol
production by clostridia. Biotechnol Bioeng 2008;101:209e28.
isomerization) can be used to produce the most desirable isomers. [21] Atsumi S, Hanai T, Liao JC. Non-fermentative pathways for synthesis of
The increased carbonyl emissions that are unavoidable when branched-chain higher alcohols as biofuels. Nature 2008;451:86e9.
burning alcohol fuels may be abated using appropriate catalytic [22] Cann AF, Liao JC. Pentanol isomer synthesis in engineered microorganisms.
Appl Microbiol Biotechnol 2010;85:893e9.
after treatment technologies. Displacing higher carbon number [23] Zhang K, Sawaya MR, Eisenberg DS, Liao JC. Expanding metabolism for
hydrocarbons with lower carbon number alcohols having similar biosynthesis of nonnatural alcohols. Proc Natl Acad Sci USA 2008;105:
physical properties can reduce soot and particulate matter emis- 20653e8.
[24] Bernton H, Kovarik W, Sklar S, Griffin B, Woolsey RJ. The forbidden fuel.
sions. Thus, this combustion chemistry review suggests that al- Bison Books; 2010.
cohols are suitable alternative fuels for a wide range of [25] Ricardo HR. The high-speed internal-combustion engine. London Glasgow:
combustion applications. A principal contribution of this review Blackie Son Ltd; 1931.
[26] Farrell AE, Plevin RJ, Turner BT, Jones AD, Hare MO, Kammen DM. Ethanol
article is presented in Box 2, wherein directions for future com- can contribute to energy and environmental goals. Science 2006;311:
bustion chemistry research are suggested to enable the clean and 506e8.
efficient utilization of alcohol fuels in future internal combustion [27] Goldemberg J. Ethanol for a sustainable energy future. Science 2007;315:
808e10.
engines.
[28] Hill J, Nelson E, Tilman D, Polasky S, Tiffany D. Environmental, economic, and
energetic costs and benefits of biodiesel and ethanol biofuels. Proc Natl Acad
Sci 2006;103:11206e10.
Acknowledgments [29] Lapola DM, Schaldach R, Alcamo J, Bondeau A, Koch J, Koelking C, et al. In-
direct land-use changes can overcome carbon savings from biofuels in Brazil.
SMS is thankful for support from Clean Combustion Research Proc Natl Acad Sci 2010;107:3388e93.
[30] Koh LP, Ghazoul J. Biofuels, biodiversity, and people: understanding the
Center at KAUST. PO acknowledges financial support within the conflicts and finding opportunities. Biol Conserv 2008;141:2450e60.
DLR Center-of-Excellence "Alternative Fuels". NH is supported by [31] Larson ED. A review of life-cycle analysis studies on liquid biofuel systems
the U.S. Department of Energy, Office of Basic Energy Sciences in for the transport sector. Energy Sustain Dev 2006;10:109e26.
[32] MacLean HL, Lave LB, Lankey R, Joshi S. A life-cycle comparison of alternative
part under the Energy Frontier Research Center for Combustion
automobile fuels. J Air Waste Manag Assoc 2000;50:1769e79.
Science (Grant No. DE-SC0001198). Sandia is a multi-program [33] Spatari S, Zhang Y, MacLean HL. Life cycle assessment of switchgrass- and
laboratory operated by Sandia Corporation, a Lockheed Martin corn stover-derived ethanol-fueled automobiles. Environ Sci Technol
2005;39:9750e8.
Company, for the National Nuclear Security Administration under
[34] Gautam M, Martin II DW. Combustion characteristics of higher-alcohol/
contract DE-AC04-94-AL85000. KKH is grateful for a sabbatical gasoline blends. Proc Inst Mech Eng A 2000;214:497e511.
period, including a research stay with the Clean Combustion [35] Saxena S, Bedoya ID. Fundamental phenomena affecting low temperature
Research Center and SMS at KAUST, which greatly facilitated the combustion and HCCI engines, high load limits and strategies for extending
these limits. Prog Energy Combust Sci 2013;39:457e88.
collaboration in the conception phase of this article. [36] Yao M, Zheng Z, Liu H. Progress and recent trends in homogenous
charge compression ignition (HCCI). Prog Energy Combust Sci 2009;35:
398e437.
Appendix A. Supplementary material [37] Christensen E, Yanowitz J, Ratcliff M, McCormick RL. Renewable oxygenate
blending effects on gasoline properties. Energy Fuels 2011;25:4723e33.
[38] Amer A, Babiker H, Chang J, Kalghatgi G, Adomeit P, Brassat A, et al. Fuel
Supplementary data relates to this article can be found at http://
effects on knock in a highly boosted direct injection spark ignition engine.
dx.doi.org/10.1016/j.pecs.2014.04.003. SAE Int J Fuels Lubr 2012;5:1048e65.
[39] Hunwartzen I. Modification of CFR test engine unit to determine octane
numbers of pure alcohols and gasoline-alcohol blends. SAE technical paper
References series; 1982:1e6.
[40] Bauer K. Klopffestigkeitsbestimmung (ROZ und MOZ) von Alkoholen und
[1] International Energy Agency. World energy outlook; 2010. Alkoholmischkraftstoffen in CFR-Prüfmotoren: Abschlußbericht. DGMK
[2] Eurostat European Commission. Energy, transport and environment in- 1981;260:1e60.
dicators. 2012 ed. Belgium: European Union; 2012. [41] Anderson JE, DiCicco DM, Ginder JM, Kramer U, Leone TG, Raney-Pablo HE,
[3] Boundy B, Diegel SW, Wright L, Davis SC. Biomass energy data book. 4th ed. et al. High octane number ethanolegasoline blends: quantifying the poten-
Oak Ridge National Laboratory; 2011. tial benefits in the United States. Fuel 2012;97:585e94.
[4] Demirbas A. Progress and recent trends in biofuels. Prog Energy Combust Sci [42] Hamadi AS. Selective additives for improvement of gasoline octane number.
2007;33:1e18. Tikrit J Eng Sci (TJES) 2010;17:22e35.
[5] Luque R, Herrero-Davila L, Campelo JM, Clark JH, Hidalgo JM, Luna D, et al. [43] Graboski MS. An analysis of alternatives for unleaded petrol additives for
Biofuels: a technological perspective. Energy Environ Sci 2008;1:542e64. South Africa. Priv Ind Rep; 2003:1e63.
96 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

[44] Giakoumis EG, Rakopoulos CD, Dimaratos AM, Rakopoulos DC. Exhaust [75] Grosjean D, Grosjean E, Gertler AW. On-road emissions of carbonyls from
emissions with ethanol or n-butanol diesel fuel blends during transient light-duty and heavy-duty vehicles. Environ Sci Technol 2001;35:45e53.
operation: a review. Renew Sustain Energy Rev 2013;17:170e90. [76] Pang X, Mu Y, Yuan J, He H. Carbonyls emission from ethanol-blended gas-
[45] Wallner T, Miers SA, McConnell S. A comparison of ethanol and butanol as oline and biodiesel-ethanol-diesel used in engines. Atmos Environ 2008;42:
oxygenates using a direct-injection, spark-ignition engine. J Eng Gas Tur- 1349e58.
bines Power 2009;131:032802. [77] Armas O, García-Contreras R, Ramos Á. Pollutant emissions from engine
[46] Miller GL, Smith JL, Workman JP. Engine performance using butanol fuel starting with ethanol and butanol diesel blends. Fuel Process Technol
blends. Trans ASAE 1981;24:538e40. 2012;100:63e72.
[47] Gautam M, Martin II DW, Carder D. Emissions characteristics of higher [78] He B-Q, Liu M-B, Yuan J, Zhao H. Combustion and emission characteristics of a
alcohol/gasoline blends. Proc Inst Mech Eng A 2000;214:165e82. HCCI engine fuelled with n-butanolegasoline blends. Fuel 2013;108:668e74.
[48] Yacoub Y, Bata R, Gautam M. The performance and emission characteristics [79] Jacobson MZ. Effects of ethanol (E85) versus gasoline vehicles on cancer and
of C1eC5 alcohol-gasoline blends with matched oxygen content in a single- mortality in the United States. Environ Sci Technol 2007;41:4150e7.
cylinder spark ignition engine. Proc Inst Mech Eng A 1998;212:1363e79. [80] Dec J. A conceptual model of DI diesel combustion based on laser-sheet
[49] Alasfour FN. Butanol-a single cylinder engine study: engine performance. Int imaging. SAE technical paper series; 1997:1e30.
J Energy Res 1997;21:21e30. [81] Musculus MPB, Miles PC, Pickett LM. Conceptual models for partially pre-
[50] Szwaja S, Naber JD. Combustion of n-butanol in a spark-ignition IC engine. mixed low-temperature diesel combustion. Prog Energy Combust Sci
Fuel 2010;89:1573e82. 2013;39:246e83.
[51] Gu X, Huang Z, Cai J, Gong J, Wu X, Lee C-F. Emission characteristics of a [82] Westbrook CK, Mizobuchi Y, Poinsot TJ, Smith PJ, Warnatz J. Computational
spark-ignition engine fuelled with gasoline-n-butanol blends in combination combustion. Proc Combust Inst 2005;30:125e57.
with EGR. Fuel 2012;93:611e7. [83] Westbrook CK, Pitz WJ, Curran HJ. Chemical kinetic modeling study of the
[52] Tornatore C, Marchitto L, Valentino G, Corcione FE, Merola SS. Optical di- effects of oxygenated hydrocarbons on soot emissions from diesel engines.
agnostics of the combustion process in a PFI SI boosted engine fueled with J Phys Chem A 2006;110:6912e22.
butanolegasoline blend. Energy 2012;45:277e87. [84] Westbrook CK. Biofuels combustion. Annu Rev Phys Chem 2013;64:201e19.
[53] Yao M, Wang H, Zheng Z, Yue Y. Experimental study of n-butanol additive [85] Westbrook CK. Chemical kinetics of hydrocarbon ignition in practical com-
and multi-injection on HD diesel engine performance and emissions. Fuel bustion systems. Proc Combust Inst 2000;28:1563e77.
2010;89:2191e201. [86] Mehl M, Chen J-Y, Pitz WJ, Sarathy SM, Westbrook CK. An approach for
[54] Lujaji F, Kristóf L, Bereczky A, Mbarawa M. Experimental investigation of fuel formulating surrogates for gasoline with application towards a reduced sur-
properties, engine performance, combustion and emissions of blends con- rogate mechanism for CFD engine modeling. Energy Fuel 2011;25:5215e23.
taining croton oil, butanol, and diesel on a CI engine. Fuel 2011;90:505e10. [87] Sarathy SM, Javed T, Karsenty F, Heufer A, Wang W, Park S, et al.
[55] Rakopoulos CD, Dimaratos AM, Giakoumis EG, Rakopoulos DC. Investigating A comprehensive combustion chemistry study of 2,5-dimethylhexane.
the emissions during acceleration of a turbocharged diesel engine operating Combust Flame 2014;161:1444e59.
with bio-diesel or n-butanol diesel fuel blends. Energy 2010;35:5173e84. [88] Kalghatgi GT. Fuel/engine interactions. Warrendale, Pennsylvania, USA: SAE
[56] Rakopoulos DC, Rakopoulos CD, Hountalas DT, Kakaras EC, Giakoumis EG, International; 2014.
Papagiannakis RG. Investigation of the performance and emissions of bus [89] Haas FM, Ramcharan A, Dryer FL. Relative reactivities of the isomeric
engine operating on butanol/diesel fuel blends. Fuel 2010;89:2781e90. butanols and ehanol in an ignition quality tester. Energy Fuels 2011;25:
[57] Rakopoulos DC, Rakopoulos CD, Giakoumis EG, Dimaratos AM, Kyritsis DC. 3909e16.
Effects of butanolediesel fuel blends on the performance and emissions of a [90] Vuilleumier D, Kozarac D, Mehl M, Saxena S, Pitz WJ, Dibble RW, et al. In-
high-speed DI diesel engine. Energy Convers Manage 2010;51:1989e97. termediate temperature heat release in an HCCI engine fueled by ethanol/n-
[58] Dogan O. The influence of n-butanol/diesel fuel blends utilization on a small heptane mixtures: an experimental and modeling study. Combust Flame
diesel engine performance and emissions. Fuel 2011;90:2467e72. 2013;161:680e95.
[59] Saisirirat P, Togbé C, Chanchaona S, Foucher F, Mounaim-Rousselle C, [91] Lu X, Hou Y, Zu L, Huang Z. Experimental study on the auto-ignition and
Dagaut P. Auto-ignition and combustion characteristics in HCCI and JSR using combustion characteristics in the homogeneous charge compression ignition
1-butanol/n-heptane and ethanol/n-heptane blends. Proc Combust Inst (HCCI) combustion operation with ethanol/n-heptane blend fuels by port
2011;33:3007e14. injection. Fuel 2006;85:2622e31.
[60] Saisirirat P, Foucher F, Chanchaona S, Mounaïm-Rousselle C. Spectroscopic [92] Lu X, Ji L, Zu L, Hou Y, Huang C, Huang Z. Experimental study and chemical
measurements of low-temperature heat release for homogeneous combus- analysis of n-heptane homogeneous charge compression ignition combus-
tion compression ignition (HCCI) n-heptane/alcohol mixture combustion. tion with port injection of reaction inhibitors. Combust Flame 2007;149:
Energy Fuels 2010;24:5404e9. 261e70.
[61] Lebedevas S, Lebedeva G, Sendzikiene E, Makareviciene V. Investigation of [93] Yates A, Bell A, Swarts A. Insights relating to the autoignition characteristics
the performance and emission characteristics of biodiesel fuel containing of alcohol fuels. Fuel 2010;89:83e93.
butanol under the conditions of diesel engine operation. Energy Fuels [94] Kamio J, Kurotani T, Kuzuoka K, Kubo Y, Taniguchi H, Hashimoto K. Study on
2010;24:4503e9. HCCI-SI combustion using fuels ethanol containing. SAE technical paper se-
[62] Mehta RN, Chakraborty M, Mahanta P, Parikh PA. Evaluation of fuel prop- ries; 2007:1e12.
erties of butanolbiodieseldiesel blends and their impact on engine per- [95] Andrae JC, Head RA. HCCI experiments with gasoline surrogate fuels
formance and emissions. Ind Eng Chem Res 2010;49:7660e5. modeled by a semidetailed chemical kinetic model. Combust Flame
[63] Zhang Y, Boehman AL. Oxidation of 1-butanol and a mixture of n-heptane/1- 2009;156:842e51.
butanol in a motored engine. Combust Flame 2010;157:1816e24. [96] Saisirirat P, Foucher F, Chanchaona S, Mounaim-Rousselle C. Effects of
[64] Alasfour FN. NOx emission from a spark ignition engine using 30% iso- ethanol, n-butanolen-heptane blended on low temperature heat release and
butanol-gasoline blend: part 1 e preheating inlet air. Appl Therm Eng HRR phasing in diesel-HCCI. SAE technical paper series; 2009:1e11.
1998;18:245e56. [97] Yang Y, Dec J, Dronniou N, Simmons B. Characteristics of isopentanol as a fuel
[65] Alasfour FN. NOx Emission from a spark ignition engine using 30% iso-butanol- for HCCI engines. SAE Int J Fuels Lubr 2010;3:725e41.
gasoline blend: part 2 e ignition timing. Appl Therm Eng 1998;18:609e18. [98] Battin-Leclerc F. Detailed chemical kinetic models for the low-temperature
[66] Alasfour FN. The effect of using 30% iso-butanol-gasoline blend on hydro- combustion of hydrocarbons with application to gasoline and diesel fuel
carbon emissions from a spark-ignition engine. Energy Sources 1999;21: surrogates. Prog Energy Combust Sci 2008;34:440e98.
379e94. [99] Zádor J, Taatjes CA, Fernandes RX. Kinetics of elementary reactions in low-
[67] Irimescu A. Performance and fuel conversion efficiency of a spark ignition temperature autoignition chemistry. Prog Energy Combust Sci 2010;37:
engine fueled with iso-butanol. Appl Energy 2012;96:477e83. 371e421.
[68] Karabektas M, Hosoz M. Performance and emission characteristics of a diesel [100] Petersen EL. Interpreting endwall and sidewall measurements in shock-tube
engine using isobutanolediesel fuel blends. Renew Energy 2009;34:1554e9. ignition studies. Combust Sci Technol 2009;181:1123e44.
[69] Al-Hasan MI, Al-Momany M. The effect of iso-butanol-diesel blends on en- [101] Hanson RK, Pang GA, Chakraborty S, Ren W, Wang S, Davidson DF. Con-
gine performance. Transport 2008;23:306e10. strained reaction volume approach for studying chemical kinetics behind
[70] Niass T, Amer AA, Xu W, Vogel SR, Krebber-Hortmann K, Adomeit P, et al. reflected shock waves. Combust Flame 2013;160:1550e8.
Butanol blending e a promising approach to enhance the thermodynamic [102] Chaos M, Dryer FL. Chemical-kinetic modeling of ignition delay: consider-
potential of gasoline e part 1. SAE Int J Fuels Lubr 2012;5:265e73. ations in interpreting shock tube data. Int J Chem Kinet 2010;42:143e50.
[71] Vancoillie J, Demuynck J, Sileghem L, Van De Ginste M, Verhelst S, Brabant L, [103] Heufer KA, Bugler J, Curran HJ. A comparison of longer alkane and alcohol
et al. The potential of methanol as a fuel for flex-fuel and dedicated spark- ignition including new experimental results for n-pentanol and n-hexanol.
ignition engines. Appl Energy 2013;102:140e9. Proc Combust Inst 2013;34:511e8.
[72] Kohse-Höinghaus K, Oßwald P, Cool TA, Kasper T, Hansen N, Qi F, et al. [104] Weber BW, Kumar K, Zhang Y, Sung C-J. Autoignition of n-butanol at
Biofuel combustion chemistry: from ethanol to biodiesel. Angew Chem Int elevated pressure and low-to-intermediate temperature. Combust Flame
Ed 2010;49:3572e97. 2011;158:809e19.
[73] Seinfeld JH, Pandis SN. Atmospheric chemistry and physics: from air pollu- [105] Weber BW, Sung C-J. Comparative autoignition trends in butanol isomers at
tion to climate change. Hoboken, New Jersey: John Wiley & Sons; 2006. elevated pressure. Energy Fuels 2013;27:1688e98.
[74] Grosjean E, Grosjean D, Fraser MP, Cass GR. Air quality model evaluation data [106] Mittal G, Sung C-J. A rapid compression machine for chemical kinetics
for organics. 2. C1eC14 carbonyls in Los Angeles air. Environ Sci Technol studies at elevated pressures and temperatures. Combust Sci Technol
1996;30:2687e703. 2007;179:497e530.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 97

[107] Glassman I, Yetter RA. Combustion. 4th ed. California: Elsevier; 2008. [139] Aniolek KW, Wilk RD. Preflame oxidation characteristics of methanol. Energy
[108] Stranic I, Chase DP, Harmon JT, Yang S, Davidson DF, Hanson RK. Shock tube Fuels 1995;9:395e405.
measurements of ignition delay times for the butanol isomers. Combust [140] Held TJ, Dryer FL. An experimental and computational study of methanol
Flame 2012;159:516e27. oxidation in the intermediate-and high-temperature regimes. Proc Combust
[109] Ranzi E, Frassoldati A, Grana R, Cuoci A, Faravelli T, Kelley AP, et al. Hierarchical Inst 1994;25:901e8.
and comparative kinetic modeling of laminar flame speeds of hydrocarbon and [141] Li SC, Williams FA. Experimental and numerical studies of two-stage
oxygenated fuels. Prog Energy Combust Sci 2012;38:468e501. methanol flames. Proc Combust Inst 1996;26:1017e24.
[110] Veloo PS, Wang YL, Egolfopoulos FN, Westbrook CK. A comparative experi- [142] Fieweger K, Blumenthal R, Adomeit G. Self-ignition of S.I. engine model fuels: a
mental and computational study of methanol, ethanol, and n-butanol flames. shock tube investigation at high pressure. Combust Flame 1997;109:599e619.
Combust Flame 2010;157:1989e2004. [143] Alzueta MU, Bilbao R, Finestra M. Methanol oxidation and its interaction
[111] Stranic I, Pyun SH, Davidson DF, Hanson RK. Multi-species measurements in with nitric oxide. Energy Fuels 2001;15:724e9.
1-butanol pyrolysis behind reflected shock waves. Combust Flame 2012;159: [144] Saeed K, Stone CR. Measurements of the laminar burning velocity for mix-
3242e50. tures of methanol and air from a constant-volume vessel using a multizone
[112] Stranic I, Pyun SH, Davidson DF, Hanson RK. Multi-species measurements in model. Combust Flame 2004;139:152e66.
2-butanol and i-butanol pyrolysis behind reflected shock waves. Combust [145] Liao SY, Jiang DM, Huang Z, Zeng K. Characterization of laminar premixed
Flame 2013;160:1012e9. methanoleair flames. Fuel 2006;85:1346e53.
[113] Hansen N, Miller JA, Taatjes CA, Wang J, Cool TA, Law ME, et al. Photoioni- [146] Dayma G, Ali KH, Dagaut P. Experimental and detailed kinetic modeling
zation mass spectrometric studies and modeling of fuel-rich allene and study of the high pressure oxidation of methanol sensitized by nitric oxide
propyne flames. Proc Combust Inst 2007;31:1157e64. and nitrogen dioxide. Proc Combust Inst 2007;31:411e8.
[114] Schenk M, Leon L, Moshammer K, Oßwald P, Zeuch T, Seidel L, et al. Detailed [147] Rasmussen CL, Wassard KH, Dam-Johansen K, Glarborg P. Methanol oxida-
mass spectrometric and modeling study of isomeric butene flames. Combust tion in a flow reactor: Implications for the branching ratio of the CH3OHþOH
Flame 2013;160:487e503. reaction. Int J Chem Kinet 2008;40:423e41.
[115] Qi F. Combustion chemistry probed by synchrotron VUV photoionization [148] Zhang Z, Huang Z, Wang X, Xiang J, Wang X, Miao H. Measurements of
mass spectrometry. Proc Combust Inst 2013;34:33e63. laminar burning velocities and Markstein lengths for methanoleairenitro-
[116] Biordi JC. Molecular beam mass spectrometry for studying the fundamental gen mixtures at elevated pressures and temperatures. Combust Flame
chemistry of flames. Prog Energy Combust Sci 1977;3:151e73. 2008;155:358e68.
[117] Korobeinichev OP, Ilyin SB, Shvartsberg VM. The destruction chemistry of [149] Togbé C, Ahmed AM, Dagaut P. Experimental and modeling study of the
organophosphorus compounds in flames e I: quantitative determination of kinetics of oxidation of methanolgasoline surrogate mixtures (M85 sur-
final phosphorus-containing species in hydrogeneoxygen flames. Combust rogate) in a jet-stirred reactor. Energy Fuels 2009;23:1936e41.
Flame 1999;118:718e26. [150] Noorani EK, Akih-Kumgeh B, Bergthorson JM. Comparative high temperature
[118] Cool TA, McIlroy A, Qi F, Westmoreland PR, Poisson L, Peterka DS, et al. shock tube ignition of C1C4 primary alcohols. Energy Fuels 2010;24:5834e43.
Photoionization mass spectrometer for studies of flame chemistry with a [151] Kumar K, Sung C-J. Autoignition of methanol: experiments and computa-
synchrotron light source. Rev Sci Instrum 2005;76:094102. tions. Int J Chem Kinet 2011;43:175e84.
[119] Qi F, Yang R, Yang B, Huang C, Wei L, Wang J, et al. Isomeric identification of [152] Vancoillie J, Christensen M, Nilsson EJK, Verhelst S, Konnov AA. Temperature
polycyclic aromatic hydrocarbons formed in combustion with tunable vac- dependence of the laminar burning velocity of methanol flames. Energy
uum ultraviolet photoionization. Rev Sci Instrum 2006;77:084101. Fuels 2012;26:1557e64.
[120] Oßwald P, Hemberger P, Bierkandt T, Akyildiz E, Köhler M, Bodi A, et al. In [153] Aranda V, Christensen JM, Alzueta MU, Glarborg P, Gersen S, Gao Y, et al.
situ flame chemistry tracing by imaging photoelectron photoion coincidence Experimental and kinetic modeling study of methanol ignition and oxidation
spectroscopy. Rev Sci Instrum 2014;85:025101. at high pressure. Int J Chem Kinet 2013;45:283e94.
[121] Harper MR, Van Geem KM, Pyl SP, Marin GB, Green WH. Comprehensive [154] Beeckmann J, Cai L, Pitsch H. Experimental investigation of the laminar
reaction mechanism for n-butanol pyrolysis and combustion. Combust Flame burning velocities of methanol, ethanol, n-propanol, and n-butanol at high
2011;158:16e41. pressure. Fuel 2014;117:340e50.
[122] Cooke DF, Dodson MG, Williams A. A shock-tube study of the ignition of [155] Rotzoll G. High-temperature pyrolysis of ethanol. J Anal Appl Pyrolysis
methanol and ethanol with oxygen. Combust Flame 1971;16:233e6. 1985;9:43e52.
[123] Bowman CT. A shock-tube investigation of the high-temperature oxidation [156] Dunphy MP, Simmie JM. High-temperature oxidation of ethanol. Part 1.
of methanol. Combust Flame 1975;25:343e54. Ignition delays in shock-waves. J Chem Soc Faraday Trans 1991;87:1691e6.
[124] Aronowitz D, Santoro RJ, Dryer FL, Glassman I. Kinetics of the oxidation of [157] Curran HJ, Dunphy MP, Simmie JM, Westbrook CK, Pitz WJ. Shock tube
methanol: experimental results semi-global modeling and mechanistic ignition of ethanol, isobutene and MTBE: experiments and modeling. Proc
concepts. Proc Combust Inst 1979;17:633e44. Combust Inst 1992;24:769e76.
[125] Singh S, Grosshandler W, Malte PC, Crain Jr RW. Oxides of nitrogen formed in [158] Dagaut P, Boettner JC, Cathonnet M. Kinetic modeling of ethanol pyrolysis
high-intensity methanol combustion. Proc Combust Inst 1979;17:689e99. and combustion. J Chim Phys Pcb 1992;89:867e84.
[126] Natarajan K, Bhaskaran KA. An experimental and analytical study of meth- [159] Egolfopoulos FN, Du DX, Law CK. A study on ethanol oxidation kinetics in
anol ignition behind shock waves. Combust Flame 1981;43:35e49. laminar premixed flames, flow reactors, and shock tubes. Proc Combust Inst
[127] Vandooren J, Van Tiggelen PJ. Experimental investigation of methanol 1992;24:833e41.
oxidation in flames: mechanisms and rate constants of elementary steps. [160] Norton TS, Dryer FL. An experimental and modeling study of ethanol
Proc Combust Inst 1981;18:473e83. oxidation kinetics in an atmospheric pressure flow reactor. Int J Chem Kinet
[128] Tsuboi T, Hashimoto K. Shock tube study on homogeneous thermal oxidation 1992;24:319e44.
of methanol. Combust Flame 1981;42:61e76. [161] Li J, Kazakov A, Dryer LF. Ethanol pyrolysis experiments in a variable pres-
[129] Gülder ÖL. Laminar burning velocities of methanol, ethanol and isooctanee sure flow reactor. Int J Chem Kinet 2001;33:859e67.
air mixtures. Proc Combust Inst 1982;19:275e81. [162] Alzueta MU, Hernández JM. Ethanol oxidation and its interaction with nitric
[130] Metghalchi M, Keck JC. Burning velocities of mixtures of air with methanol, oxide. Energy Fuels 2002;16:166e71.
isooctane, and indolene at high pressure and temperature. Combust Flame [163] Benvenutti LH, Marques CST, Bertran CA. Chemiluminescent emission data
1982;48:191e210. for kinetic modeling of ethanol combustion. Combust Sci Technol 2004;177:
[131] Pauwels JF, Carlier M, Devolder P, Sochet LR. Experimental and numerical 1e26.
analysis of a low pressure stoichiometric methanol-air flame. Combust Sci [164] Hidaka Y, Wakamatsu H, Moriyama M, Koike T, Yasunaga K. Shock-tube
Technol 1989;64:97e117. study of ethanol pyrolysis. In: Shock Waves. Berlin Heidelberg: Springer;
[132] Norton TS, Dryer FL. Some new observations on methanol oxidation chem- 2005: 651e6.
istry. Combust Sci Technol 1989;63:107e29. [165] Ergut A, Granata S, Jordan J, Carlson J, Howard JB, Richter H, et al. PAH for-
[133] Norton TS, Dryer FL. The flow reactor oxidation of C1eC4 alcohols and MTBE. mation in one-dimensional premixed fuel-rich atmospheric pressure ethyl-
Proc Combust Inst 1991;23:179e85. benzene and ethyl alcohol flames. Combust Flame 2006;144:757e72.
[134] Bradley D, Dixon-Lewis G, El-Din Habik S, Kwa LK, El-Sherif S. Laminar flame [166] Kasper TS, Oßwald P, Kamphus M, Kohse-Höinghaus K. Ethanol flame
structure and burning velocities of premixed methanoleair. Combust Flame structure investigated by molecular beam mass spectrometry. Combust
1991;85:105e20. Flame 2007;150:220e31.
[135] Cribb PH, Dove JE, Yamazaki S. A kinetic study of the pyrolysis of methanol [167] Liao SY, Jiang DM, Huang ZH, Zeng K. Determination of the laminar burning
using shock tube and computer simulation techniques. Combust Flame velocities for mixtures of ethanol and air at elevated temperatures. Appl
1992;88:169e85. Therm Eng 2007;27:374e80.
[136] Cribb PH, Dove JE, Yamazaki S. A kinetic study of the oxidation of methanol [168] Saxena P, Williams FA. Numerical and experimental studies of ethanol
using shock tube and computer simulation techniques. Combust Flame flames. Proc Combust Inst 2007;31:1149e56.
1992;88:186e200. [169] Dagaut P, Togbé C. Experimental and modeling study of the kinetics of
[137] Egolfopoulos FN, Du DX, Law CK. A comprehensive study of methanol ki- oxidation of ethanolgasoline surrogate mixtures (E85 surrogate) in a jet-
netics in freely-propagating and burner-stabilized flames, flow and static stirred reactor. Energy Fuels 2008;22:3499e505.
reactors, and shock tubes. Combust Sci Technol 1992;83:33e75. [170] Leplat N, Seydi A, Vandooren J. An experimental study of the structure of a
[138] Lee D, Hochgreb S, Keck JC. Autoignition of alcohols and ethers in a rapid stoichiometric ethanol/oxygen/argon flame. Combust Sci Technol 2008;180:
compression machine. SAE Technical paper 932755; 1993:1e8. 519e32.
98 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

[171] Wang J, Yang B, Cool TA, Hansen N, Kasper T. Near-threshold absolute [199] Esarte C, Abian M, Millera A, Bilbao R, Alzueta MU. Gas and soot products
photoionization cross-sections of some reaction intermediates in combus- formed in the pyrolysis of acetylene mixed with methanol, ethanol, iso-
tion. Int J Mass Spectrom 2008;269:210e20. propanol or n-butanol. Energy 2012;43:37e46.
[172] Bradley D, Lawes M, Mansour MS. Explosion bomb measurements of [200] Man X, Tang C, Zhang J, Zhang Y, Pan L, Huang Z, et al. An experimental and
ethanoleair laminar gaseous flame characteristics at pressures up to 1.4MPa. kinetic modeling study of n-propanol and i-propanol ignition at high tem-
Combust Flame 2009;156:1462e70. peratures. Combust Flame 2014;161:644e56.
[173] Cancino LR, Fikri M, Oliveira AAM, Schulz C. Measurement and chemical [201] Barnard JA. The pyrolysis of n-butanol. Trans Faraday Soc 1957;53:1423e30.
kinetics modeling of shock-induced ignition of ethanolair mixtures. Energy [202] Barnard JA. The pyrolysis of tert-butanol. Trans Faraday Soc 1959;55:947e51.
Fuels 2010;24:2830e40. [203] Tsang W. Thermal decomposition of some tert-butyl compounds at elevated
[174] Heufer KA, Olivier H. Determination of ignition delay times of different hy- temperatures. J Chem Phys 1964;40:1498e505.
drocarbons in a new high pressure shock tube. Shock Waves 2010;20:307e [204] Lewis D, Keil M, Michael S, Sarr M. Gas phase thermal decomposition of tert-
16. butyl alcohol. J Am Chem Soc 1974;96:4398e404.
[175] Eisazadeh-Far K, Moghaddas A, Al-Mulki J, Metghalchi H. Laminar burning [205] Choudhury T, Lin M, Lin C, Sanders W. Thermal-decomposition of tert-butyl
speeds of ethanol/air/diluent mixtures. Proc Combust Inst 2011;33:1021e7. alcohol in shock-waves. Combust Sci Technol 1990;71:219e32.
[176] Konnov AA, Meuwissen RJ, de Goey LPH. The temperature dependence of the [206] McEnally C, Pfefferle L. Fuel decomposition and hydrocarbon growth pro-
laminar burning velocity of ethanol flames. Proc Combust Inst 2011;33: cesses for oxygenated hydrocarbons: butyl alcohols. Proc Combust Inst
1011e9. 2005;30:1363e70.
[177] Leplat N, Dagaut P, Togbé C, Vandooren J. Numerical and experimental study [207] Moss JT, Berkowitz AM, Oehlschlaeger MA, Biet J, Warth V, Glaude P-A, et al.
of ethanol combustion and oxidation in laminar premixed flames and in jet- An experimental and kinetic modeling study of the oxidation of the four
stirred reactor. Combust Flame 2011;158:705e25. isomers of butanol. J Phys Chem A 2008;112:10843e55.
[178] van Lipzig JPJ, Nilsson EJK, de Goey LPH, Konnov AA. Laminar burning ve- [208] Dagaut P, Sarathy SM, Thomson MJ. A chemical kinetic study of n-butanol
locities of n-heptane, iso-octane, ethanol and their binary and tertiary oxidation at elevated pressure in a jet-stirred reactor. Proc Combust Inst
mixtures. Fuel 2011;90:2773e81. 2009;32:229e37.
[179] Xu H, Yao C, Yuan T, Zhang K, Guo H. Measurements and modeling study of [209] Dagaut P, Togbé C. Experimental and modeling study of the kinetics of
intermediates in ethanol and dimethyl ether low-pressure premixed flames oxidation of butanolen-heptane mixtures in a jet-stirred reactor. Energy
using synchrotron photoionization. Combust Flame 2011;158:1673e81. Fuels 2009;23:3527e35.
[180] Broustail G, Seers P, Halter F, Moréac G, Mounaim-Rousselle C. Experimental [210] Gu X, Huang Z, Li Q, Tang C. Measurements of laminar burning velocities and
determination of laminar burning velocity for butanol and ethanol iso- Markstein lengths of n-butanolair premixed mixtures at elevated tem-
octane blends. Fuel 2011;90:1e6. peratures and pressures. Energy Fuels 2009;23:4900e7.
[181] Lee C, Vranckx S, Heufer KA, Khomik SV, Uygun Y, Olivier H, et al. On the [211] Sarathy SM, Thomson MJ, Togbé C, Dagaut P, Halter F, Mounaim-Rousselle C.
chemical kinetics of ethanol oxidation: shock tube, rapid compression ma- An experimental and kinetic modeling study of n-butanol combustion.
chine and detailed modeling study. Z Phys Chem 2012;226:1e28. Combust Flame 2009;156:852e64.
[182] Varea E, Modica V, Vandel A, Renou B. Measurement of laminar burning [212] Black G, Curran HJ, Pichon S, Simmie JM, Zhukov V. Bio-butanol: combustion
velocity and Markstein length relative to fresh gases using a new post- properties and detailed chemical kinetic model. Combust Flame 2010;157:
processing procedure: application to laminar spherical flames for methane, 363e73.
ethanol and isooctane/air mixtures. Combust Flame 2012;159:577e90. [213] Grana R, Frassoldati A, Faravelli T, Niemann U, Ranzi E, Seiser R, et al. An
[183] Broustail G, Halter F, Seers P, Moréac G. Experimental determination of experimental and kinetic modeling study of combustion of isomers of
laminar burning velocity for butanol/iso-octane and ethanol/iso-octane butanol. Combust Flame 2010;157:2137e54.
blends for different initial pressures. Fuel 2013;106:310e7. [214] Gu X, Huang Z, Wu S, Li Q. Laminar burning velocities and flame instabilities
[184] Tran L-S, Glaude P-A, Fournet R, Battin-Leclerc F. Experimental and modeling of butanol isomerseair mixtures. Combust Flame 2010;157:2318e25.
study of premixed laminar flames of ethanol and methane. Energy Fuels [215] Togbé C, Mzé-Ahmed A, Dagaut P. Kinetics of oxidation of 2-butanol and
2013;27:2226e45. isobutanol in a jet-stirred reactor: experimental study and modeling
[185] Varea E, Modica V, Renou B, Boukhalfa AM. Pressure effects on laminar investigation. Energy Fuels 2010;24:5244e56.
burning velocities and Markstein lengths for isooctaneeethanoleair mix- [216] Van Geem K, Pyl S, Marin G, Harper M, Green W. Accurate high-temperature
tures. Proc Combust Inst 2013;34:735e44. reaction networks for alternative fuels: butanol isomers. Ind Eng Chem Res
[186] Herrmann F, Jochim B, Oßwald P, Cai L, Pitsch H, Kohse-Höinghaus K. 2010;21:2e16.
Experimental and numerical low-temperature oxidation study of ethanol [217] Gu X, Li Q, Huang Z, Zhang N. Measurement of laminar flame speeds and
and dimethyl ether. Combust Flame 2014;116:384e97. flame stability analysis of tert-butanoleair mixtures at elevated pressures.
[187] Mittal G, Burke SM, Davies VA, Parajuli B, Metcalfe WK, Curran HJ. Auto- Energy Convers Manage 2011;52:3137e46.
ignition of ethanol in a rapid compression machine. Combust Flame [218] Hansen N, Harper MR, Green WH. High-temperature oxidation chemistry of
2014;161:1164e71. n-butanol e experiments in low-pressure premixed flames and detailed
[188] Knorsch T, Zackel A, Mamaikin D, Zigan L, Wensing M. Comparison of kinetic modeling. Phys Chem Chem Phys 2011;13:20262e74.
different gasoline alternative fuels in terms of laminar burning velocity at [219] Heufer KA, Fernandes RX, Olivier H, Beeckmann J, Röhl O, Peters N. Shock
increased Gas temperatures and exhaust gas recirculation rates. Energy Fuels tube investigations of ignition delays of n-butanol at elevated pressures
2014;28:1446e52. between 770 and 1250K. Proc Combust Inst 2011;33:359e66.
[189] Barnard JA. The pyrolysis of isopropanol. Trans Faraday Soc 1960;56:72e9. [220] Karwat DMA, Wagnon SW, Teini PD, Wooldridge MS. On the chemical ki-
[190] Barnard JA, Hughes HWD. The pyrolysis of n-propanol. Trans Faraday Soc netics of n-butanol: ignition and speciation studies. J Phys Chem A 2011;115:
1960;56:64e71. 4909e21.
[191] Sinha A, Thomson MJ. The chemical structures of opposed flow diffusion [221] Oßwald P, Güldenberg H, Kohse-Höinghaus K. Combustion of butanol iso-
flames of C3 oxygenated hydrocarbons (isopropanol, dimethoxy methane, and mers e a detailed molecular beam mass spectrometry investigation of their
dimethyl carbonate) and their mixtures. Combust Flame 2004;136:548e56. flame chemistry. Combust Flame 2011;158:2e15.
[192] Li Y, Wei L, Tian Z, Yang B, Wang J, Zhang T, et al. A comprehensive exper- [222] Veloo P, Egolfopoulos F. Flame propagation of butanol isomers/air mixtures.
imental study of low-pressure premixed C3-oxygenated hydrocarbon flames Proc Combust Inst 2011;33:987e93.
with tunable synchrotron photoionization. Combust Flame 2008;152:336e [223] Vranckx S, Heufer KA, Lee C, Olivier H, Schill L, Kopp WA, et al. Role of peroxy
59. chemistry in the high-pressure ignition of n-butanol e experiments and
[193] Johnson MV, Goldsborough SS, Serinyel Z, O’Toole P, Larkin E, O’Malley G, detailed kinetic modelling. Combust Flame 2011;158:1444e55.
et al. A shock tube study of n- and iso-propanol ignition. Energy Fuels [224] Cai J, Zhang L, Yang J, Li Y, Zhao L, Qi F. Experimental and kinetic
2009;23:5886e98. modeling study of tert-butanol combustion at low pressure. Energy
[194] Kasper T, Oßwald P, Struckmeier U, Kohse-Höinghaus K, Taatjes CA, Wang J, 2012;43:94e102.
et al. Combustion chemistry of the propanol isomers e investigated by [225] Cai J, Zhang L, Zhang F, Wang Z, Cheng Z, Yuan W, et al. Experimental and
electron ionization and VUV-photoionization molecular-beam mass spec- kinetic modeling study of n-butanol pyrolysis and combustion. Energy Fuel
trometry. Combust Flame 2009;156:1181e201. 2012;26:5550e68.
[195] Frassoldati A, Cuoci A, Faravelli T, Niemann U, Ranzi E, Seiser R, et al. An [226] Karwat DMA, Wagnon SW, Wooldridge MS, Westbrook CK. On the com-
experimental and kinetic modeling study of n-propanol and iso-propanol bustion chemistry of n-heptane and n-butanol blends. J Phys Chem A
combustion. Combust Flame 2010;157:2e16. 2012;116:12406e21.
[196] Akih-Kumgeh B, Bergthorson JM. Ignition of C3 oxygenated hydrocarbons [227] Lefkowitz JK, Heyne JS, Won SH, Dooley S, Kim HH, Haas FM, et al. A chemical
and chemical kinetic modeling of propanal oxidation. Combust Flame kinetic study of tertiary-butanol in a flow reactor and a counterflow diffu-
2011;158:1877e89. sion flame. Combust Flame 2012;159:968e78.
[197] Galmiche B, Togbé C, Dagaut P, Halter F, Foucher F. Experimental and [228] Liu W, Kelley A, Law C. Nonpremixed ignition, laminar flame propagation,
detailed kinetic modeling study of the oxidation of 1-propanol in a pres- and mechanism reduction of n-butanol, iso-butanol, and methyl butanoate.
surized jet-stirred reactor (JSR) and a combustion bomb. Energy Fuels Proc Combust Inst 2011;33:995e1002.
2011;25:2013e21. [229] Van Geem KM, Cuoci A, Frassoldati A, Pyl SP, Marin GB, Ranzi E. An exper-
[198] Veloo PS, Egolfopoulos FN. Studies of n-propanol, iso-propanol, and propane imental and kinetic modeling study of pyrolysis and combustion of acetone-
flames. Combust Flame 2011;158:501e10. butanoleethanol (ABE) mixtures. Combust Sci Technol 2012;184:942e55.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 99

[230] Yasunaga K, Mikajiri T, Sarathy SM, Koike T. A shock tube and chemical ki- [260] Sarathy SM, Vranckx S, Yasunaga K, Mehl M, Oßwald P, Metcalfe WK, et al.
netic modeling study of the pyrolysis and oxidation of butanols. Combust A comprehensive chemical kinetic combustion model for the four butanol
Flame 2012;159:2009e27. isomers. Combust Flame 2012;159:2028e55.
[231] Chung GA, Akih-Kumgeh B, Watson GMG, Bergthorson JM. NOx formation [261] Yi F, Axelbaum RL. Stability of spray combustion for water/alcohols mixtures
and flame velocity profiles of iso- and n-isomers of butane and butanol. Proc in oxygen-enriched air. Proc Combust Inst 2013;34:1697e704.
Combust Inst 2013;34:831e8. [262] Chen G, Yu W, Jiang X, Huang Z, Wang Z, Cheng Z. Experimental and
[232] Hansen N, Merchant SS, Harper MR, Green WH. The predictive capability of modeling study on the influences of methanol on premixed fuel-rich n-
an automatically generated combustion chemistry mechanism: chemical heptane flames. Fuel 2013;103:467e72.
structures of premixed iso-butanol flames. Combust Flame 2013;160:2343e [263] Kohse-Höinghaus K, Oßwald P, Struckmeier U, Kasper T, Hansen N,
51. Taatjes CA, et al. The influence of ethanol addition on premixed fuel-rich
[233] Merchant SS, Zanoelo EF, Speth RL, Harper MR, Van Geem KM, Green WH. propeneeoxygeneargon flames. Proc Combust Inst 2007;31:1119e27.
Combustion and pyrolysis of iso-butanol: Experimental and chemical kinetic [264] Hamins A, Seshadri K. The influence of alcohols on the combustion of hy-
modeling study. Combust Flame 2013;160:1907e29. drocarbon fuels in diffusion flames. Combust Flame 1986;64:43e54.
[234] Zhang J, Pan L, Mo J, Gong J, Huang Z, Law CK. A shock tube and kinetic [265] Chen G, Yu W, Fu J, Mo J, Huang Z, Yang J, et al. Experimental and modeling
modeling study of n-butanal oxidation. Combust Flame 2013;160:1541e9. study of the effects of adding oxygenated fuels to premixed n-heptane
[235] Cai J, Yuan W, Ye L, Cheng Z, Wang Y, Zhang L, et al. Experimental and kinetic flames. Combust Flame 2012;159:2324e35.
modeling study of 2-butanol pyrolysis and combustion. Combust Flame [266] Wu J, Song KH, Litzinger T, Lee S-Y, Santoro R, Linevsky M, et al. Reduction of
2013;160:1939e57. PAH and soot in premixed ethyleneeair flames by addition of ethanol.
[236] Wu F, Law CK. An experimental and mechanistic study on the laminar flame Combust Flame 2006;144:675e87.
speed, Markstein length and flame chemistry of the butanol isomers. [267] Bell KM, Tipper CFH. The effect of surface on the slow combustion of
Combust Flame 2013;160:2744e56. methanol. Trans Faraday Soc 1957;53:982e90.
[237] Camacho J, Lieb S, Wang H. Evolution of size distribution of nascent soot in n- [268] Tester JW, Webley PA, Holgate HR. Revised global kinetic measurements
and i-butanol flames. Proc Combust Inst 2013;34:1853e60. of methanol oxidation in supercritical water. Ind Eng Chem Res 1993;32:
[238] Vasu SS, Sarathy SM. On the high-temperature combustion of n-butanol: shock 236e9.
tube data and an improved kinetic model. Energy Fuels 2013;27:7072e80. [269] Vogel F, Blanchard J, Marrone PA, Rice SF. Critical review of kinetic data for
[239] Yang Z, Qian Y, Yang X, Wang Y, Wang Y, Huang Z, et al. Autoignition of n- the oxidation of methanol in supercritical water. J Supercrit Fluids 2005;34:
butanol/n-heptane blend fuels in a rapid compression machine under low- 249e86.
to-medium temperature ranges. Energy Fuels 2013;27:7800e8. [270] Dagaut P, Cathonnet M, Boettner J-C. Chemical kinetic modeling of the
[240] Zhu Y, Davidson DF, Hanson RK. 1-Butanol ignition delay times at low supercritical-water oxidation of methanol. J Supercrit Fluids 1996;9:33e
temperatures: An application of the constrained-reaction-volume strategy. 42.
Combust Flame 2014;161:634e43. [271] Brock EE, Oshima Y, Savage PE. Kinetics and mechanism of methanol
[241] Cai J, Yuan W, Ye L, Cheng Z, Wang Y, Dong W, et al. Experimental and kinetic oxidation in supercritical water. J Phys Chem 1996;100:15834e42.
modeling study of i-butanol pyrolysis and combustion. Combust Flame; [272] Cathonnet M, Boettner JC, James H. Experimental-study and simulation of
2014. http://dx.doi.org/10.1016/j.combustflame.2014.02.004. methanol pyrolysis. J Chim Phys PCB 1979;76:183e9.
[242] Jin H, Wang Y, Zhang K, Guo H, Qi F. An experimental study on the formation [273] Cathonnet M, Boettner JC, James H. Study of methanol oxidation and self
of polycyclic aromatic hydrocarbons in laminar coflow non-premixed ignition in the temperature-range 500e600-degrees-c. J Chim Phys PCB
methane/air flames doped with four isomeric butanols. Proc Combust Inst 1982;79:475e8.
2013;34:779e86. [274] Fieweger K, Ciezki H, Adomeit G. Comparison of shock-tube ignition character-
[243] Pan L, Zhang Y, Tian Z, Yang F, Huang Z. Experimental and kinetic study on istics of various fuel-air mixtures at high pressures. In: Shock Waves. Berlin
ignition delay times of iso-butanol. Energy Fuels 2014;28:2160e9. Heidelberg: Springer; 1995: 161e6.
[244] Togbé C, Dagaut P, Mzé-Ahmed A, Diévart P, Halter F, Foucher F. Experi- [275] Dunphy MP, Simmie JM. High-temperature oxidation of ethanol. Part 1 e
mental and detailed kinetic modeling study of 1-hexanol oxidation in a ignition delays in shock waves. J Chem Soc Faraday Trans 1991;87:1691e6.
pressurized jet-stirred reactor and a combustion bomb. Energy Fuels [276] Wang J, Struckmeier U, Yang B, Cool TA, Oßwald P, Kohse-Höinghaus K, et al.
2010;24:5859e75. Isomer-specific influences on the composition of reaction intermediates in
[245] Dayma G, Togbé C, Dagaut P. Experimental and detailed kinetic modeling dimethyl ether/propene and ethanol/propene flame. J Phys Chem A
study of isoamyl alcohol (isopentanol) oxidation in a jet-stirred reactor at 2008;112:9255e65.
elevated pressure. Energy Fuels 2011;25:4986e98. [277] Gao J, Zhao D-Q, Wang X-H, Jiang L-Q, Yang H-L, Yuan T, et al. Comparative
[246] Togbé C, Halter F, Foucher F, Mounaim-Rousselle C, Dagaut P. Experimental study of dimethyl ether and ethanol premixed laminar flames at low pres-
and detailed kinetic modeling study of 1-pentanol oxidation in a JSR and sure. Acta Phys Chim Sin 2010;26:23e8.
combustion in a bomb. Proc Combust Inst 2011;33:367e74. [278] Xu H, Yao C, Yuan T, Zhang K, Guo H. Measurements and modeling study of
[247] Mzé-Ahmed A, Hadj-Ali K, Diévart P, Dagaut P. Kinetics of oxidation of a intermediates in ethanol and dimethyl ether low pressure premixed flames
reformulated jet fuel (1-hexanol/jet A-1) in a jet-stirred reactor: experi- using synchrotron photo ionization. Combust Flame 2011;158:1673e81.
mental and modeling study. Combust Sci Technol 2012;184:1039e50. [279] Cancino LR, Fikri M, Oliveira AAM, Schulz C. Ignition delay times of ethanol-
[248] Heufer KA, Sarathy SM, Curran HJ, Davis AC, Westbrook CK, Pitz WJ. containing multi-component gasoline surrogates: shock-tube experiments
Detailed kinetic modeling study of n-pentanol oxidation. Energy Fuel and detailed modeling. Fuel 2011;90:1238e44.
2012;11:6678e85. [280] Simmie JM. Detailed chemical kinetic models for the combustion of hydro-
[249] Tsujimura T, Pitz WJ, Gillespie F, Curran HJ, Weber BW, Zhang Y, et al. carbon fuels. Prog Energy Combust Sci 2003;29:599e634.
Development of isopentanol reaction mechanism reproducing autoignition [281] Kee RJ, Coltrin ME, Glarborg P. Chemically reacting flow. Wiley-Interscience;
character at high and low temperatures. Energy Fuel 2012;8:4871e86. 2005.
[250] Tang C, Wei L, Man X, Zhang J, Huang Z, Law CK. High temperature ignition [282] CHEMKIN-PRO 15112. Reaction design; 2011. San Diego, CA.
delay times of C5 primary alcohols. Combust Flame 2013;160:520e9. [283] CANTERA Version: 2.1.1.www.cantera.org; 2014.
[251] Li Q, Hu E, Cheng Y, Huang Z. Measurements of laminar flame speeds and [284] Goodwin D. Cantera: object-oriented software for reacting flows. Technical
flame instability analysis of 2-methyl-1-butanol-air mixtures. Fuel report. California Institute of Technology; 2005.
2013;112:263e71. [285] Cuoci A, Frassoldati A, Faravelli T, Ranzi E. A computational tool for the
[252] Sarathy SM, Park S, Weber BW, Wang W, Veloo PS, Davis AC, et al. detailed kinetic modeling of laminar flames: application to C2H4/CH4 coflow
A comprehensive experimental and modeling study if iso-pentanol com- flames. Combust Flame 2013;160:870e86.
bustion. Combust Flame 2013;160:2712e28. [286] Cuoci A, Frassoldati A, Faravelli T, Ranzi E. Formation of soot and nitrogen
[253] Yeung C, Thomson MJ. Experimental and kinetic modeling study of 1- oxides in unsteady counterflow diffusion flames. Combust Flame 2009;156:
hexanol combustion in an opposed-flow diffusion flame. Proc Combust 2010e22.
Inst 2013;34:795e802. [287] Computational Modelling Cambridge Ltd. Kinetics: the chemical kinetics
[254] Fletcher CJM. The Thermal decomposition of methyl alcohol. Proc R Soc Lond model builder http://www.cmclinnovations.com/; 2013.
A 1934;147:119e28. [288] Deutschmann O, Tischer S, Kleditzsch S, Janardhanan VM, Correa C,
[255] Fort R, Hinshelwood CN. Further investigations on the kinetics of gaseous Chatterjee D, et al. DETCHEM software package; 2013.
oxidation reactions. Proc R Soc Lond A 1930;129:284e99. [289] Frenklach M, Wang H, Rabinowitz MJ. Optimization and analysis of large
[256] Bone WA, Gardner JB. Comparative studies of the slow combustion of chemical kinetic mechanisms using the solution mapping method e com-
methane, methyl alcohol, formaldehyde, and formic acid. Proc R Soc Lond A bustion of methane. Prog Energy Combust Sci 1992;18:47e73.
1936;154:297e328. [290] Benson SW, Cohen N. Current status of group additivityIn ACS symposium
[257] Schultz RF, Kistiakowsky GB. The Thermal decomposition of tertiary butyl series, vol. 677. American Chemical Society; 2009. pp. 20e46.
and tertiary amyl alcohols. Homogeneous unimolecular reactions. J Am [291] Benson SW. Thermochemical kinetics. 2nd ed. New York: Wiley; 1976.
Chem Soc 1934;56:395e8. [292] Ritter E, Bozzelli J. Therm - thermodynamic property estimation for Gas-
[258] Hinshelwood CN, Lennard-Jones JE, Travers M, Polnayi M, Zener C, Bowen EJ, Phase radicals and molecules. Int J Chem Kinet 1991;23:767e78.
et al. Discussion on energy distribution in molecules in relation to chemical [293] Goldsmith CF, Magoon GR. A database of small molecule thermochemistry
reactions. Opening address. Proc R Soc Lond A; 1934:239e71. for combustion. J Phys Chem A 2012;116:9033e57.
[259] Semenov NN. Chemical kinetics and chain reactions. The Clarendon Press; [294] Wang H, Frenklach M. Transport-properties of polycyclic aromatic-
1935. hydrocarbons for flame modeling. Combust Flame 1994;96:163e70.
100 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

[295] Tee L, Gotoh S, Stewart W. Molecular parameters for normal fluids e [328] Veloo PS, Dagaut P, Togbé C, Dayma G, Sarathy SM, Westbrook CK, et al.
Lennard-Jones 12-6 potential. Ind Eng Chem Fund 1966;3:356e63. Experimental and modeling study of the oxidation of n- and iso-butanal.
[296] Holley A, You X, Dames E, WANG H, Egolfopoulos F. Sensitivity of propa- Combust Flame 2013;160:1609e26.
gation and extinction of large hydrocarbon flames to fuel diffusion. Proc [329] da Silva G, Kim C-H, Bozzelli JW. Thermodynamic properties (enthalpy, bond
Combust Inst 2009;32:1157e63. energy, entropy, and heat capacity) and internal rotor potentials of vinyl
[297] Brown NJ, Bastien LAJ, Price PN. Transport properties for combustion alcohol, methyl vinyl ether, and their corresponding radicals. J Phys Chem A
modeling. Prog Energ Combust Sci 2011;37:565e82. 2006;110:7925e34.
[298] Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, [330] da Silva G, Bozzelli J. Role of the [a]-hydroxyethylperoxy radical in the re-
et al. Gaussian 09. Wallingford, CT: Gaussian Inc; 2009. actions of acetaldehyde and vinyl alcohol with HO2. Chem Phys Lett
[299] Montgomery JA, Frisch MJ, Ochterski JW, Petersson GA. A complete basis set 2009;483:25e9.
model chemistry. VII. Use of the minimum population localization method. [331] Huynh LK, Zhang HR, Zhang S, Eddings E. Kinetics of enol formation from
J Chem Phys 2000;112:6532e42. reaction of OH with propene. J Phys Chem A 2009;113:3177e85.
[300] Curtiss LA, Raghavachari K, Redfern PC, Rassolov V, Pople JA. Gaussian-3 (G3) [332] Zádor J, Jasper AW, Miller JA. The reaction between propene and hydroxyl.
theory for molecules containing first and second-row atoms. J Chem Phys Phys Chem Chem Phys 2009;11:11040e53.
1998;109:7764e76. [333] da Silva G. Carboxylic acid catalyzed keto-enol tautomerizations in the Gas
[301] Curtiss LA, Redfern PC, Raghavachari K. Gaussian-4 theory using reduced phase. Angew Chem Int Ed 2010;49:7523e5.
order perturbation theory. J Chem Phys 2007;127:124105. [334] Rao H-B, Zeng X-Y, He H, Li Z-R. Theoretical investigations on removal re-
[302] Carstensen HH, Dean AM. Development of Detailed Kinetic Models For The actions of ethenol by H Atom. J Phys Chem A 2011;115:1602e8.
Thermal Conversion of Biomass Via First Principle Methods And Rate Esti- [335] Altarawneh M, Al-Muhtaseb AH, Dlugogorski BZ, Kennedy EM, Mackie JC.
mation Rules. In: Computational modeling in lignocellulosic biofuel pro- Rate constants for hydrogen abstraction reactions by the hydroperoxyl
duction. ACE Symposium Series. New York: Springer; 2010:201e43. radical from methanol, ethenol, acetaldehyde, toluene, and phenol. J Comp
[303] Westbrook CK, Dryer FL. Chemical kinetic modeling of hydrocarbon com- Chem 2011;32:1725e33.
bustion. Prog Energy Combust Sci 1984;10:1e57. [336] Cavalli F, Geiger H, Barnes I, Becker K. FTIR kinetic, product, and modeling
[304] Basevich VY. Chemical kinetics in the combustion processes: a detailed ki- study of the OH-initiated oxidation of 1-butanol in air. Environ Sci Technol
netics mechanism and its implementation. Prog Energy Combust Sci 2002;36:1263e70.
1987;13:199e248. [337] Hurley M, Wallington T, Laursen L, Javadi M, Nielsen O, Yamanaka T, et al.
[305] von Elbe G, Lewis B. The mechanism of the combustion of hydrocarbons. Atmospheric chemistry of n-butanol: kinetics, mechanisms, and products of
Proc Combust Inst 1948;1e2:169e74. Cl atom and OH radical initiated oxidation in the presence and absence of
[306] von Elbe G. Chemical kinetics of hydrocarbon combustion. Proc Combust Inst NOx. J Phys Chem A 2009;113:7011e20.
1955;5:79e85. [338] Andersen VF, Wallington TJ, Nielsen OJ. Atmospheric chemistry of i-butanol.
[307] Lewis B, von Elbe G. Combustion, flames and explosions of gases. Orlando J Phys Chem A 2010;114:12462e9.
and London: Academic Press Inc; 1987. [339] Zádor J, Fernandes RX, Georgievskii Y, Meloni G, Taatjes CA, Miller JA. The
[308] Miller JA, Bowman CT. Mechanism and modeling of nitrogen chemistry in reaction of hydroxyethyl radicals with O2: a theoretical analysis and
combustion. Prog Energy Combust Sci 1989;15:287e338. experimental product study. Proc Combust Inst 2009;32:271e7.
[309] Westbrook CK, Dryer FL. Comprehensive mechanism for methanol oxidation. [340] da Silva G, Bozzelli JW, Liang L, Farrell JT. Ethanol oxidation: kinetics of the a-
Combust Sci Technol 1979;20:125e40. hydroxyethyl radicalþO2 reaction. J Phys Chem A 2009;113:8923e33.
[310] Marinov NM. A detailed chemical kinetic model for high temperature [341] Jasper A, Klippenstein S, Hardin L. Theoretical rate coefficients for the re-
ethanol oxidation. Int J Chem Kinet 1999;31:183e220. action of methyl radical with hydroperoxyl radical and for methylhy-
[311] Tran L-S, Sirjean B, Glaude P-A, Fournet R, Battin-Leclerc F. Progress in droperoxide decomposition. Proc Combust Inst 2009;32:279e86.
detailed kinetic modeling of the combustion of oxygenated components of [342] Ray DJM, Waddington DJ. Gas phase oxidation of alkenes e part II. The
biofuels. Energy 2012;43:4e18. oxidation of 2-methylbutene-2 and 2,3-dimethylbutene-2. Combust Flame
[312] Held TJ, Dryer FL. A comprehensive mechanism for methanol oxidation. Int J 1973;20:327e34.
Chem Kinet 1998;30:805e30. [343] Sway MI, Waddington DJ. Reactions of oxygenated radicals in the gas phase.
[313] Li J, Zhao Z, Kazakov A, Chaos M, Dryer FL, Scire JJ. A comprehensive kinetic Part 12. The reactions of isopropylperoxyl radicals and alkenes. J Chem Soc
mechanism for CO, CH2O, and CH3OH combustion. Int J Chem Kinet 2007;39: Perkin Trans 1983;2:139e43.
109e36. [344] Sun H, Bozzelli J, Law C. Thermochemical and kinetic analysis on the re-
[314] Metcalfe WK, Burke SM, Ahmed SS, Curran HJ. A hierarchical and compar- actions of O2 with products from OH addition to isobutene, 2-hydroxy-1, 1-
ative kinetic modeling study of C1C2 hydrocarbon and oxygenated fuels. dimethylethyl, and 2-hydroxy-2-methylpropyl radicals: HO2 formation from
Int J Chem Kinet 2013;45:638e75. oxidation of neopentane, part II. J Phys Chem A 2007;111:4974e86.
[315] Battin-Leclerc F, Blurock E, Bounaceur R, Fournet R, Glaude P-A, Herbinet O, [345] Welz O, Zádor J, Savee JD, Ng MY, Meloni G, Fernandes RX, et al. Low-tem-
et al. Towards cleaner combustion engines through groundbreaking detailed perature combustion chemistry of biofuels: pathways in the initial low-
chemical kinetic models. Chem Soc Rev 2011;40:4762e82. temperature (550 Ke750 K) oxidation chemistry of isopentanol. Phys
[316] Touchard S, Fournet R, Glaude PA, Warth V, Battin-Leclerc F, Vanhove G, et al. Chem Chem Phys 2012;14:3112.
Modeling of the oxidation of large alkenes at low temperature. Proc Combust [346] Welz O, Zádor J, Savee JD, Sheps L, Osborn DL, Taatjes CA. Low-temperature
Inst 2005;30:1073e81. combustion chemistry of n-butanol: principal oxidation pathways of
[317] Pierucci S, Ranzi E. A review of features in current automatic generation hydroxybutyl radicals. J Phys Chem A 2013;117:11983e2001.
software for hydrocarbon oxidation mechanisms. Comp Chem Eng 2008;32: [347] Quelch GE, Gallo MM, Shen M, Xie Y, Schaefer HF, Moncrieff D. Aspects of the
805e26. reaction mechanism of ethane combustion. 2. Nature of the intramolecular
[318] Frassoldati A, Grana R, Faravelli T, Ranzi E, Oßwald P, Kohse-Höinghaus K. hydrogen transfer. J Am Chem Soc 1994;116:4953e62.
Detailed kinetic modeling of the combustion of the four butanol [348] Jorand F, Heiss A, Sahetchian K, Kerhoas L, Einhorn J. Identification of an
isomers in premixed low-pressure flames. Combust Flame 2012;159: unexpected peroxide formed by successive isomerization reactions of
2295e311. the n-butoxy radical in oxygen. J Chem Soc Faraday Trans 1996;92:
[319] Curran HJ, Gaffuri P, Pitz WJ, Westbrook CK. A comprehensive modeling 4167e71.
study of n-heptane oxidation. Combust Flame 1998;114:149e77. [349] Heiss A, Sahetchian K. Isomerization reactions of the n-C4H9O and n-
[320] Curran HJ, Gaffuri P, Pitz WJ, Westbrook CK. A comprehensive modeling OOC4H8OH radicals in oxygen. Int J Chem Kinet 1996;28:531e44.
study of iso-octane oxidation. Combust Flame 2002;129:253e80. [350] Perrin O, Heiss A, Doumenc F, Sahetchian K. Homogeneous and heteroge-
[321] Warnatz J. Hydrocarbon oxidation at high temperatures. Ber Bunsenges Phys neous reactions of the n-C5H11O, n-C5H10OH and OOC5H10OH radicals in
Chem 1983;87:1008e22. oxygen. Analytical steady state solution by use of the Laplace transform.
[322] Dove JE, Warnatz J. Calculation of burning velocity and flame structure in J Chem Soc Faraday Trans 1998;94:2323e35.
methanol e air mixtures. Ber Bunsenges Phys Chem 1983;87:1040e4. [351] Perrin O, Heiss A, Sahetchian K. Determination of the isomerization rate
[323] Kéromnès A, Metcalfe WK, Heufer KA, Donohoe N, Das AK, Sung C-J, et al. An constant HOCH2CH2CH2CH(OO)CH3/ HOC$HCH2CH2CH(OOH)CH3. Impor-
experimental and detailed chemical kinetic modeling study of hydrogen and tance of intramolecular hydroperoxy isomerization in tropospheric chem-
syngas mixture oxidation at elevated pressures. Combust Flame 2013;160: istry. Int J Chem Kinet 1998;30:875e87.
995e1011. [352] Baulch DL, Cobos CJ, Cox RA, Esser C, Frank P, Just Th, et al. Evaluated kinetic
[324] Healy D, Kalitan DM, Aul CJ, Petersen EL, Bourque G, Curran HJ. Oxidation of data for combustion modeling: supplement II. J Phys Chem Ref Data
C1C5 alkane quinternary natural gas mixtures at high pressures. Energy 2005;34:757e1397.
Fuels 2010;24:1521e8. [353] Jiménez E, Gilles MK, Ravishankara AR. Kinetics of the reactions of the hy-
[325] Moc J, Simmie JM, Curran HJ. The elimination of water from a conforma- droxyl radical with CH3OH and C2H5OH between 235 and 360 K.
tionally complex alcohol: a computational study of the gas phase dehydra- J Photochem Photobiol Chem 2003;157:237e45.
tion of n-butanol. J Mol Struc 2009;928:149e57. [354] Dillon TJ, Hölscher D, Sivakumaran V, Horowitz A, Crowley JN. Kinetics of the
[326] Zhou C-W, Simmie JM, Curran HJ. Rate constants for hydrogen abstraction by reactions of HO with methanol (210e351 K) and with ethanol (216e368 K).
H2O from n-butanol. Int J Chem Kinet 2012;44:155e64. Phys Chem Chem Phys 2005;7:349e55.
[327] Veloo PS, Dagaut P, Togbé C, Dayma G, Sarathy SM, Westbrook CK, et al. Jet- [355] Srinivasan NK, Su MC, Michael JV. High-temperature rate constants for
stirred reactor and flame studies of propanal oxidation. Proc Combust Inst CH3OHþKr/products, OHþCH3OH/products, OHþ(CH3) 2CO/CH2COCH3þ
2013;34:599e606. H2O, and OHþCH3/CH2þH2O. J Phys Chem A 2007;111:3951e8.
S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102 101

[356] Lu C-W, Chou S-L, Lee Y-P, Xu S, Xu ZF, Lin MC. Experimental and theoretical kinetic isotope effects, transition structure, reaction path, and mechanism.
studies of rate coefficients for the reaction O(3P)þCH3OH at high tempera- J Phys Chem A 2004;108:11554e8.
tures. J Chem Phys 2005;122:244314e22. [387] Rosado-Reyes CM, Tsang W. Shock tube study on the thermal decomposition
[357] Carr SA, Blitz MA, Seakins PW. Product branching fractions for the reaction of of n-butanol. J Phys Chem A 2012;116:9825e31.
O(3P) atoms with methanol and ethanol. Chem Phys Lett 2011;511:207e12. [388] Rosado-Reyes CM, Tsang W. Shock tube studies on the decomposition of 2-
[358] Park J, Xu ZF, Xu K, Lin MC. Kinetics for the reactions of phenyl with butanol. J Phys Chem A 2012;116:9599e606.
methanol and ethanol: comparison of theory and experiment. Proc Combust [389] Rosado-Reyes C, Tsang W. Bond cleavage during isobutanol Thermal
Inst 2012;34:473e82. decomposition and the breaking of CeC bonds in alcohols at high temper-
[359] Lu K-W, Matsui H, Huang C-L, Raghunath P, Wang N-S, Lin MC. Shock tube atures. J Phys Chem A 2013;117:10170e7.
study on the thermal decomposition of CH3OH. J Phys Chem A 2010;114: [390] Rosado-Reyes CM, Tsang W, Alecu IM, Merchant SS, Green WH. Dehydration
5493e502. of isobutanol and the elimination of water from fuel alcohols. J Phys Chem A
[360] Lee P-F, Matsui H, Xu D-W, Wang N-S. Thermal decomposition and oxidation 2013;117:6724e36.
of CH3OH. J Phys Chem A 2013;117:525e34. [391] Wallington TJ, Dagaut P, Liu R, Kurylo MJ. Rate constants for the gas phase
[361] Meier U, Grotheer HH, Riekert G, Just T. Temperature dependence and branching reactions of OH with C5 through C7 aliphatic alcohols and ethers: predicted
ratio of the C2H5OHþOH reaction. Chem Phys Lett 1985;115:221e5. and experimental values. Int J Chem Kinet 1988;20:541e7.
[362] Greenhill PG, Ogrady BV. The rate-constant of the reaction of hydroxyl radicals [392] Ing W-C, Sheng CY, Bozzelli JW. Development of a detailed high-pressure
with methanol, ethanol and (D3)Methanol. Aust J Chem 1986;39:1775e87. reaction model for methane/methanol mixtures under pyrolytic and oxida-
[363] Wallington TJ, Kurylo MJ. The gas phase reactions of hydroxyl radicals with a tive conditions and comparison with experimental data. Fuel Proc Tech
series of aliphatic alcohols over the temperature range 240e440 K. Int J 2003;83:111e45.
Chem Kinet 1987;19:1015e23. [393] Jasper AW, Klippenstein SJ, Harding LB, Ruscic B. Kinetics of the reaction of
[364] Hess WP, Tully FP. Catalytic conversion of alcohols to alkenes by OH. Chem methyl radical with hydroxyl radical and methanol decomposition. J Phys
Phys Lett 1988;152:183e9. Chem A 2007;111:3932e50.
[365] Bott JF, Cohen N. A shock tube study of the reactions of the hydroxyl radical [394] Meana-Pañeda R, Truhlar DG, Fernández-Ramos A. High-level direct-
with several combustion species. Int J Chem Kinet 1991;23:1075e94. dynamics variational transition state theory calculations including multidi-
[366] Sivaramakrishnan R, Su M, Michael J, Klippenstein S, Harding L, Ruscic B. mensional tunneling of the thermal rate constants, branching ratios, and
Rate constants for the thermal decomposition of ethanol and its bimolecular kinetic isotope effects of the hydrogen abstraction reactions from methanol
reactions with OH and d: reflected shock tube and theoretical studies. J Phys by atomic hydrogen. J Chem Phys 2011;134:094302.
Chem A 2010;114:9425e39. [395] Xu S, Lin MC. Theoretical study on the kinetics for OH reactions with CH3OH
[367] Carr SA, Blitz MA, Seakins PW. Site-specific rate coefficients for reaction of and C2H5OH. Proc Combust Inst 2007;31:159e66.
OH with ethanol from 298 to 900 K. J Phys Chem A 2011;115:3335e45. [396] Klippenstein SJ, Harding LB, Davis MJ, Tomlin AS, Skodje RT. Uncertainty
[368] Orkin VL, Khamaganov VG, Martynova LE, Kurylo MJ. High-accuracy mea- driven theoretical kinetics studies for CH3OH ignition: HO2þCH3OH and
surements of OH reaction rate constants and IR and UV absorption spectra: O2þCH3OH. Proc Combust Inst 2011;33:351e7.
ethanol and partially fluorinated ethyl alcohols. J Phys Chem A 2011;115: [397] Alecu IM, Truhlar DG. Computational study of the reactions of methanol with
8656e68. the hydroperoxyl and methyl radicals. 1. Accurate thermochemistry and
[369] Stranic I, Pang GA, Hanson RK, Golden DM, Bowmant CT. Shock tube mea- barrier heights. J Phys Chem A 2011;115:2811e29.
surements of the rate constant for the reaction ethanol plus OH. J Phys Chem [398] Alecu IM, Truhlar DG. Computational study of the reactions of methanol with
A 2014;118:822e8. the hydroperoxyl and methyl radicals. 2. Accurate thermal rate constants.
[370] Li J, Kazakov A, Dryer FL. Experimental and numerical studies of ethanol J Phys Chem A 2011;115:14599e611.
decomposition reactions. J Phys Chem A 2004;108:7671e80. [399] Hippler H, Striebel F, Viskolcz B. A detailed experimental and theoretical
[371] Wu C-W, Matsui H, Wang N-S, Lin MC. Shock tube study on the thermal study on the decomposition of methoxy radicals. Phys Chem Chem Phys
decomposition of ethanol. J Phys Chem A 2011;115:8086e92. 2001;3:2450e8.
[372] Dunlop JR, Tully FP. Catalytic dehydration of alcohols by hydroxyl: 2- [400] Grotheer HH, Riekert G, Walter D, Just T. Direct study of the reactions of
propanol: an intermediate case. J Phys Chem 1993;97:6457e64. CH2OH and CH3CHOH radicals with O(3P) atoms. Chem Phys Lett 1988;148:
[373] Yujing M, Mellouki A. Temperature dependence for the rate constants of the 530e6.
reaction of OH radicals with selected alcohols. Chem Phys Lett 2001;333:63e8. [401] Gersen S, Mokhov AV, Darmeveil JH, Levinsky HB, Glarborg P. Ignition-pro-
[374] Rajakumar B, McCabe DC, Talukdar RK, Ravishankara AR. Rate coefficients for moting effect of NO2 on methane, ethane and methane/ethane mixtures in a
the reactions of OH with n-propanol and iso-propanol between 237 and rapid compression machine. Proc Combust Inst 2011;33:433e40.
376 K. Int J Chem Kinet 2010;42:10e24. [402] Gersen, Anikin, Mokhov, Levinsky. Ignition properties of methane/hydrogen
[375] Orkin VL, Khamaganov VG, Kurylo MJ. High accuracy measurements of OH mixtures in a rapid compression machine. Int J Hydrogen Energy 2008;33:
reaction rate constants and IR absorption spectra: substituted 2-propanols. 1957e64.
J Phys Chem A 2012;116:6188e98. [403] Frassoldati A, Cuoci A, Faravelli T, Ranzi E. Kinetic modeling of the oxidation
[376] Wallington TJ, Dagaut P, Liu R. Gas-phase reactions of hydroxyl radicals with of ethanol and gasoline surrogate mixtures. Combust Sci Technol 2010;182:
the fuel additives methyl tert-butyl ether and tert-butyl alcohol over the 653e67.
temperature range 240e440 K. Environ Sci Technol 1988;22:842e4. [404] Haas FM, Chaos M, Dryer FL. Low and intermediate temperature oxidation of
[377] Teton S, Mellouki A, LeBras G, Sidebottom H. Rate constants for reactions of ethanol and ethanol–PRF blends: An experimental and modeling study.
OH radicals with a series of asymmetrical ethers and tert-butyl alcohol. Int J Combust Flame 2009;156:2346e50.
Chem Kinet 1996;28:291e7. [405] Herzler J, Manion JA, Tsang W. Single-pulse shock tube study of the
[378] Mellouki A, Oussar F, Lun X, Chakir A. Kinetics of the reactions of the OH decomposition of tetraethoxysilane and related compounds. J Phys Chem A
radical with 2-methyl-1-propanol, 3-methyl-1-butanol and 3-methyl-2- 1997;101:5500e8.
butanol between 241 and 373 K. Phys Chem Chem Phys 2004;6:2951e5. [406] Tsang W. Energy transfer effects during the multichannel decomposition of
[379] Jiménez E, Lanza B, Garzón A, Ballesteros B, Albaladejo J. Atmospheric ethanol. Int J Chem Kinet 2004;36:456e65.
degradation of 2-butanol, 2-methyl-2-butanol, and 2,3-dimethyl-2-butanol: [407] Park J, Zhu RS, Lin MC. Thermal decomposition of ethanol. I. Ab initio mo-
OH kinetics and UV absorption cross sections. J Phys Chem A 2005;109: lecular orbital/RiceeRamspergereKasseleMarcus prediction of rate constant
10903e9. and product branching ratios. J Chem Phys 2002;117:3224e31.
[380] Vasu SS, Davidson DF, Hanson RK, Golden DM. Measurements of the reaction [408] Aders WK, Wagner HG. Untersuchungen zur Reaktion von Wasser-
of OH with n-butanol at high-temperatures. Chem Phys Lett 2010;497:26e9. stoffatomen mit Äthanol und tert. Butanol. Ber Bunsenges Phys Chem
[381] Pang GA, Hanson RK, Golden DM, Bowman CT. Rate constant measurements 1973;77:712e8.
for the overall reaction of OHþ1-butanol/products from 900 to 1200 K. [409] Park J, Xu ZF, Lin MC. Thermal decomposition of ethanol. II. A computational
J Phys Chem A 2012;116:2475e83. study of the kinetics and mechanism for the HþC2H5OH reaction. J Chem
[382] Pang GA, Hanson RK, Golden DM, Bowman CT. High-temperature rate con- Phys 2003;118:9990e6.
stant determination for the reaction of OH with iso-butanol. J Phys Chem A [410] Zheng J, Truhlar DG. Multi-path variational transition state theory for
2012;116:4720e5. chemical reaction rates of complex polyatomic species: ethanol þ OH re-
[383] Pang GA, Hanson RK, Golden DM, Bowman CT. Experimental determination actions. Faraday Disc 2012;157:59e88.
of the high-temperature rate constant for the reaction of OH with sec- [411] Dean AM, Bozzelli JW. Combustion chemistry of nitrogen. In: Gardiner Jr WC,
butanol. J Phys Chem A 2012;116:9607e13. editor. Gas-phase combustion chemistry. 2nd ed. New York: Springer; 2000:
[384] Stranic I, Pang GA, Hanson RK, Golden DM, Bowman CT. Shock tube mea- 125e332.
surements of the tert-butanolþOH reaction rate and the tert-C4H8OH radical [412] Xu ZF, Xu K, Lin MC. Ab initio kinetics for decomposition/isomerization re-
b-scission branching ratio using isotopic labeling. J Phys Chem A 2013;117: actions of C2H5O radicals. Chem Phys Chem 2009;10:972e82.
4777e84. [413] Senosiain JP, Klippenstein SJ, Miller JA. Reaction of ethylene with hydroxyl
[385] McGillen MR, Baasandorj M, Burkholder JB. Gas-phase rate coefficients for radicals: a theoretical study. J Phys Chem A 2006;110:6960e70.
the OH þ n-, i-, s-, and t-butanol reactions measured between 220 and [414] Dames EE. Master equation modeling of the unimolecular decompositions of
380 K: non-Arrhenius behavior and site-specific reactivity. J Phys Chem A a-hydroxyethyl (CH3CHOH) and ethoxy (CH3CH2O) radicals. Int J Chem Kinet
2013;117:4636e56. 2014;46:176e88.
[386] Kalra BL, Lewis DK, Singer SR, Raghavan AS, Baldwin JE, Hess BA. Thermal [415] Miyoshi A, Matsui H, Washida N. Reactions of hydroxyethyl radicals with
unimolecular elimination of water from tert-butyl alcohol: deuterium oxygen and nitric oxide. Chem Phys Lett 1989;160:291e4.
102 S.M. Sarathy et al. / Progress in Energy and Combustion Science 44 (2014) 40e102

[416] Grotheer HH, Riekert G, Walter D, Just T. Reactions of hydroxymethyl and [447] Yang Y, Dec J, Dronniou N, Cannella W. Boosted HCCI combustion using low-
hydroxyethyl radicals with molecular and atomic oxygen. Proc Combust Inst octane gasoline with fully premixed and partially stratified charges. SAE
1989;22:963e72. International; 2012:1e14.
[417] Ji C, Sarathy SM, Veloo PS, Westbrook CK, Egolfopoulos FN. Effects of fuel [448] McEnally CS, Pfefferle LD. Sooting tendencies of oxygenated hydrocarbons in
branching on the propagation of octane isomers flames. Combust Flame laboratory-scale flames. Environ Sci Technol 2011;45:2498e503.
2012;159:1426e36. [449] McEnally CS, Ciuparu DM, Pfefferle LD. Experimental study of fuel decom-
[418] Green EM. Fermentative production of butanol e the industrial perspective. position and hydrocarbon growth processes for practical fuel components:
Curr Opin Biotechnol 2011;22:337e43. heptanes. Combust Flame 2003;134:339e53.
[419] Dagaut P, Togbé C. Oxidation kinetics of butanol-gasoline surrogate mixtures [450] Pepiot-Desjardins P, Pitsch H, Malhotra R, Kirby S, Boehman A. Structural
in a jet-stirred reactor: experimental and modeling study. Fuel 2008;87: group analysis for soot reduction tendency of oxygenated fuels. Combust
3313e21. Flame 2008;154:191e205.
[420] Green WH, Barton PI, Bhattacharjee B, Matheu DM, Schwer DA, Song J, et al. [451] Zheng J, Meana-Pañeda R, Truhlar DG. Prediction of experimentally un-
Computer construction of detailed chemical kinetic models for gas-phase available product branching ratios for biofuel combustion: the role of
reactors. Ind Eng Chem Res 2001;40:5362e70. anharmonicity in the reaction of isobutanol with OH. J Am Chem Soc
[421] Green WH, Allen JW, Buesser BA, Ashcraft RW, Beran GJ, Class CA, et al. RMG 2014;136:5150e60.
e reaction mechanism generator v4.0.1 http://rmg.sourceforge.net/; 2013. [452] Cai L, Pitsch H. Mechanism optimization based on reaction rate rules.
[422] Zhou C-W, Klippenstein SJ, Simmie JM, Curran HJ. Theoretical kinetics for the Combust Flame 2014;161:405e15.
decomposition of iso-butanol and related reactions. Proc Combust Inst [453] Frenklach M. Transforming data into knowledgedProcess Informatics for
2013;31:501e9. combustion chemistry. Proc Combust Inst 2007;31:125e40.
[423] El-Nahas AM, Mangood AH, Takeuchi H, Taketsugu T. Thermal decomposi- [454] Russi T, Packard A, Frenklach M. Uncertainty quantification: making pre-
tion of 2-butanol as a potential nonfossil fuel: a computational study. J Phys dictions of complex reaction systems reliable. Chem Phys Lett 2010;499:1e8.
Chem A 2011;115:2837e46. [455] You X, Russi T, Packard A, Frenklach M. Optimization of combustion kinetic
[424] Ratkiewicz A, Bieniewska J, Truong TN. Kinetics of the hydrogen abstraction models on a feasible set. Proc Combust Inst 2011;33:509e16.
ROH þ H / ROH þ H2 reaction class: an application of the reaction class [456] Sheen D, Wang H. Combustion kinetic modeling using multispecies time
transition state theory. Int J Chem Kinet 2010;43:414e29. histories in shock-tube oxidation of heptane. Combust Flame 2011;158:
[425] Orme J, Curran H, Simmie J. Experimental and modeling study of methyl 645e56.
cyclohexane pyrolysis and oxidation. J Phys Chem A 2006;110:114e31. [457] Tomlin AS, Turányi T, Pilling MJ. Mathematical tools for the construction,
[426] Zhou C, Simmie J, Curran H. Rate Constants for Hydrogen-Abstraction by OH investigation and reduction of combustion mechanisms. Compr Chem Kinet
from n-Butanol. Combust Flame 2011;158:726e31. 1997;35:293e437.
[427] Seal P, Oyedepo G, Truhlar DG. Kinetics of the hydrogen atom abstraction [458] Turányi T, Nagy T, Zsély IG, Cserháti M, Varga T, Szabó BT, et al. Determi-
reactions from 1-butanol by hydroxyl radical: theory matches experiment nation of rate parameters based on both direct and indirect measurements.
and more. J Phys Chem A 2013;117:275e82. Int J Chem Kinet 2012;44:284e302.
[428] Galano A, Alvarez-Idaboy JRL, Bravo-Perez G, Ruiz-Santoyo ME. Gas phase [459] Nagy T, Turányi T. Uncertainty of Arrhenius parameters. Int J Chem Kinet
reactions of C1-C4 alcohols with the OH radical: a quantum mechanical 2011;43:359e78.
approach. Phys Chem Chem Phys 2002;4:4648e62. [460] Wang Y, Raj A, Chung SH. Combust Flame 2013;160:1667e76.
[429] El-Nahas AM, Mangood AH, El-Shereafy EE, El-Meleigy AB. Thermochemistry [461] Raj A, Prada IDC, Amer AA, Chung SH. A reaction mechanism for gasoline
and kinetics of isobutanol oxidation by the OH radical. Fuel 2013;106:431e6. surrogate fuels for large polycyclic aromatic hydrocarbons. Combust Flame
[430] Bethel HL, Atkinson R, Arey J. Kinetics and products of the reactions of 2012;159:500e15.
selected diols with the OH radical. Int J Chem Kinet 2001;33:310e6. [462] Dagaut P, Glarborg P, Alzueta M. The oxidation of hydrogen cyanide and
[431] Zheng J, Truhlar DG. Kinetics of hydrogen-transfer isomerizations of butoxyl related chemistry. Prog Energy Combust Sci 2008;34:1e46.
radicals. Phys Chem Chem Phys 2010;12:7782e93. [463] Saffaripour M, Kholghy M, Dworkin SB, Thomson MJ. A numerical and
[432] Xu X, Papajak E, Zheng J, Truhlar DG. Multi-structural variational transition experimental study of soot formation in a laminar coflow diffusion flame of a
state theory: kinetics of the 1,5-hydrogen shift isomerization of the 1- jet A-1 surrogate. Proc Combust Inst 2013;34:1057e65.
butoxyl radical including all structures and torsional anharmonicity. Phys [464] Saffaripour M, Veshkini A, Kholghy M, Thomson MJ. Experimental investi-
Chem Chem Phys 2012;14:4204e16. gation and detailed modeling of soot aggregate formation and size distri-
[433] Zhang P, Klippenstein SJ, Law CK. Ab initio kinetics for the decomposition of bution in laminar coflow diffusion flames of jet A-1, a synthetic kerosene,
hydroxybutyl and butoxy radicals of n-butanol. J Phys Chem A 2013;117: and n-decane. Combust Flame 2013;161:848e63.
1890e906. [465] Xuan Y, Blanquart G. Numerical modeling of sooting tendencies in a laminar
[434] Simmie J, Curran H. Energy barriers for the addition of H, CH3, and C2H5 to co-flow diffusion flame. Combust Flame 2013;160:1657e66.
CH2-CHX [X¼ H, CH3, OH] and for H-Atom addition to RCH-O [R¼H, CH3, [466] Bisetti F, Blanquart G, Mueller ME, Pitsch H. On the formation and early
C2H4, n-C3H7]. J Phys Chem A 2009;113:7834e45. evolution of soot in turbulent nonpremixed flames. Combust Flame
[435] Daranlot J, Bergeat A, Caralp F, Caubet P, Costes M, Forst W, et al. Gas-Phase 2012;159:317e35.
kinetics of hydroxyl radical reactions with alkenes: experiment and theory. [467] Attili A, Bisetti F, Mueller ME, Pitsch H. Formation, growth, and transport of
Chem Phys Chem 2010;11:4002e10. soot in a three-dimensional turbulent non-premixed jet flame. Combust
[436] Welz O, Savee JD, Eskola AJ, Sheps L, Osborn DL, Taatjes CA. Low-tempera- Flame 2014;161:1849e65.
ture combustion chemistry of biofuels: pathways in the low-temperature [468] Foong TM, Morganti KJ, Brear MJ, da Silva G, Yang Y, Dryer FL. The octane
(550e700K) oxidation chemistry of isobutanol and tert-butanol. Proc numbers of ethanol blended with gasoline and its surrogates. Fuel 2014;115:
Combust Inst 2012;34:493e500. 727e39.
[437] Davis SG, Law CK. Determination of and fuel structure effects on laminar flame [469] Sjöberg M, Dec JE, Hwang W. Thermodynamic and chemical effects of EGR
speeds of C1 to C8 hydrocarbons. Combust Sci Technol 1998;140:427e49. and its constituents on HCCI autoignition. SAE International; 2007:1e19.
[438] Zhao L, Ye L, Zhang F, Zhang L. Thermal decomposition of 1-pentanol and its [470] Visakhamoorthy S, Wen JZ, Sivoththaman S, Koch CR. Numerical study of a
isomers: a theoretical study. J Phys Chem A 2012;116:9238e44. butanol/heptane fuelled homogeneous charge compression ignition (HCCI)
[439] Welz O, Klippenstein SJ, Harding LB, Taatjes CA, Zádor J. Unconventional engine utilizing negative valve overlap. Appl Energy 2012;94:166e73.
peroxy chemistry in alcohol oxidation: the water elimination pathway. [471] Zuehl J, Ghandhi J, Hagen C. Fuel effects on HCCI combustion using negative
J Phys Chem Lett 2013;4:350e4. valve overlap. SAE technical paper series; 2010:1e14.
[440] Ji C, Dames E, Wang YL, Wang H, Egolfopoulos FN. Propagation and extinc- [472] Hagen LM, Olesky LM, Bohac SV, Lavoie G, Assanis D. Effects of a low octane
tion of premixed C5eC12 n-alkane flames. Combust Flame 2010;157:277e87. gasoline blended fuel on negative valve overlap enabled HCCI load limit, com-
[441] Kalghatgi GT, Bradley D. Pre-ignition and “super-knock” in turbo-charged bustion phasing and burn duration. J Eng Gas Turbine Power 2013;135:072001.
spark-ignition engines. Int J Eng Res 2012;13:399e414. [473] Dronniou N, Dec J. Investigating the development of thermal stratification
[442] Mehl M, Pitz WJ, Westbrook CK, Curran HJ. Kinetic modeling of gasoline from the near-wall regions to the bulk-gas in an HCCI engine with planar
surrogate components and mixtures under engine conditions. Proc Combust imaging thermometry. SAE International; 2012:1e29.
Inst 2011;33:193e200. [474] Sjöberg M, Dec JE. Smoothing HCCI heat-release rates using partial fuel strat-
[443] Kalghatgi GT. Auto-ignition quality of practical fuels and implications for fuel ification with two-stage ignition fuels. SAE technical paper series; 2006:1e19.
requirements of future SI and HCCI engines. SAE technical paper series; [475] Yang Y, Dec JE, Dronniou N, Sjöberg M. Tailoring HCCI heat-release rates with
2005:1e20. partial fuel stratification: comparison of two-stage and single-stage-ignition
[444] Chang J, Kalghatgi G, Amer A, Viollet Y. Enabling high efficiency direct in- fuels. Proc Combust Inst 2011;33:3047e55.
jection engine with naphtha fuel through partially premixed charge [476] Mehl M, Pitz WJ, Sarathy SM, Yang Y, Dec J. Detailed kinetic modeling of
compression ignition combustion. SAE International; 2012:1e18. conventional gasoline at highly boosted conditions and the associate inter-
[445] Chang J, Viollet Y, Amer A, Kalghatgi G. Fuel economy potential of partially mediate temperature heat release. SAE International; 2012. 1e11.
premixed compression ignition (PPCI) combustion with naphtha fuel. SAE [477] Leppard WR. The chemical origin of fuel octane sensitivity. SAE technical
International; 2013:1e14. paper series; 1990:1e17.
[446] Kalghatgi GT, Head RA. The available and required autoignition quality of [478] Mehl M, Faravelli T, Giavazzi F, Ranzi E, Scorletti P, Tardani A, et al. Detailed
gasoline-like fuels in HCCI engines at high temperatures. SAE technical paper chemistry promotes understanding of octane numbers and gasoline sensi-
series; 2004:1e17. tivity. Energy Fuels 2006;20:2391e8.

S-ar putea să vă placă și