Sunteți pe pagina 1din 115

Experimental and Kinetic Modeling Study of Ethyl Levulinate

Oxidation in a Jet-Stirred Reactor

Thesis by
Jui-Yang Wang

In Partial Fulfillment of the Requirements


For the Degree of
Master of Science in Chemical and Biological Engineering

King Abdullah University of Science and Technology


Thuwal, Kindom of Saudi Arabia

May, 2017
2

EXAMINATION COMMITTEE PAGE

The thesis of Jui-Yang Wang is approved by the examination committee.

Committee Chairperson: Professor Subram Maniam Sarathy


Committee Members: Professor Aamir Farooq, Professor Robert W.
Dibble, Professor Klaus-Viktor Peinemann
3

COPYRIGHT PAGE

© May, 2017
Jui-Yang Wang
All Rights Reserved
4

ABSTRACT

A jet-stirred reactor was designed and constructed in the Clean Combustion

Research Center (CCRC) at King Abdullah University of Science and Technology

(KAUST); was validated with n-heptane, iso-octane oxidation and cyclohexene pyrolysis.

Different configurations of the setup have been tested to achieve good agreement with

results from the literature. Test results of the reactor indicated that installation of a pumping

system at the downstream side in the experimental apparatus was necessary to avoid the

reoccurrence of reactions in the sampling probe.

Experiments in ethyl levulinate oxidation were conducted in the reactor under

several equivalence ratios, from 600 to 1000 K, 1 bar and 2 s residence time. Oxygenated

species detected included methyl vinyl ketone, levulinic acid and ethyl acrylate. Ethylene,

methane, carbon monoxide, hydrogen, oxygen and carbon dioxide were further quantified

with a gas chromatography, coupled with a flame ionization detector and a thermal

conductivity detector.

The ethyl levulinate chemical kinetic model was first developed by Dr. Stephen

Dooley, Trinity College Dublin, and simulated under the same conditions, using the

Perfect-Stirred Reactor code in Chemkin software. In comparing the simulation results

with experimental data, some discrepancies were noted; predictions of ethylene production

were not well matched. The kinetic model was improved by updating several classes of

reactions: unimolecular decomposition, H-abstraction, C-C and C-O beta-scissions of fuel

radicals. The updated model was then compared again with experimental results and good

agreement was achieved, proving that the concerted eliminated reaction is crucial for the

kinetic mechanism formulation of ethyl levulinate. In addition, primary reaction pathways


5

and sensitivity analysis were performed to describe the role of molecular structure in

combustion (800 and 1000 K for ethyl levulinate oxidation in the jet-stirred reactor).
6

ACKNOWLEDGEMENTS

I would like to thank my advisor, Prof. S. Mani Sarathy, who provided me plenty

of opportunities and ideas for this work and inspired me in different ways.

I thank my thesis committee members, including Profs. Aamir Farooq, Robert W.

Dibble, and Klaus-Viktor Peinemann, for correcting my thesis and providing valuable

comments. I would especially like to thank Dr. Zhandong for helping out setting up the

jest-stirred reactor with unlimited support, assistance, and advices.

It is a privilege to finish the thesis alongside a group of talented and passionate

scientists in Clean Combustion Research Center. A very big thank you to Nour Atef and

Samah Mohamed for their help on optimizing the kinetic model and simulation. To several

patient academic- and life-mentors: Eshan Singh, Ehson Nasir, and Bingjie Chen who

helped me a lot on experimental setup and validation, I am deeply indebted. Special thanks

to Tony Bissoonauth for improving the fluency of my thesis.

To some senior scientists Zeyad Talla, Misjudeen Raji, and Najeh Kharbatia from

Analytical Core Laboratory, I express my appreciation for their kindly technical assistance

on coupling my setup with their apparatus for species analysis.

Particularly, many thanks go to some best friends inside or outside KAUST, Joel,

Aniela, Jam, Mafer, Moe, Joao, Daniel, Zhibek, Omar, Dora, Ken, Charlene, and Robert

for being there with me for all the ups and lows and sharing their memories and experience

with me.

Finally I would like to thank my parents and my sister for their patience,

encouragement and support through my education.


7

TABLE OF CONTENTS

EXAMINATION COMMITTEE PAGE .........................................................................2


COPYRIGHT PAGE .........................................................................................................3
ABSTRACT ........................................................................................................................4
ACKNOWLEDGEMENTS ..............................................................................................6
LIST OF FIGURES ...........................................................................................................9
LIST OF TABLES ...........................................................................................................12
Chapter 1 INTRODUCTION .........................................................................................13
Chapter 2 BACKGROUND RESEARCH .....................................................................15
2.1 Introduction to Bio-derived Fuels..................................................................................... 15
2.2 Biofuel Use in Engines ....................................................................................................... 15
2.3 Bio-derived Fuel Combustion Chemistry ........................................................................ 17
2.3.1 Alcohols Chemistry ..................................................................................................... 17
2.3.2 Biodiesel Chemistry .................................................................................................... 18
2.4 From Lignocellulosic Biomass to Ethyl Levulinate ........................................................ 21
Chapter 3 JET-STIRRED REACTOR VALIDATION ...............................................25
3.1 Literature Review on Biofuels Combustion in JSR ........................................................ 25
3.2 Experimental Set-Up at KAUST ...................................................................................... 30
3.2.1 Fuel Injection System.................................................................................................. 33
3.3 Verification of Criteria for KAUST JSR ......................................................................... 36
3.4 Temperature Homogeneity Measurement ....................................................................... 38
3.5 Sampling Methods ............................................................................................................. 39
3.5.1 Downstream Sampling Design ................................................................................... 42
3.6 Analytical Techniques ....................................................................................................... 45
3.6.1 Gas Chromatography ................................................................................................. 46
3.6.2 Column Selection ........................................................................................................ 48
3.6.3 Principles of GC as a Refinery Gas Analyzer (RGA) .............................................. 49
3.6.4 Principles of GC/MS/FID/TCD.................................................................................. 52
Chapter 4 JSR VALIDATION EXPERIMENT ...........................................................58
4.1 N-heptane Oxidation.......................................................................................................... 58
4.1.1 Oxidation Experiment of n-heptane .......................................................................... 61
4.2 Cyclohexene Pyrolysis ....................................................................................................... 65
4.3 Iso-octane Oxidation .......................................................................................................... 67
Chapter 5 ETHYL LEVULINATE MODELING AND EXPERIMENT ..................69
5.1 Ethyl levulinate Oxidation Mechanism ............................................................................ 69
8

5.1.1 Kinetic Model .............................................................................................................. 69


5.1.2 Ethyl Levulinate Thermochemistry .......................................................................... 74
5.2 Ethyl Levulinate Oxidation Experiment .......................................................................... 75
5.3 Kinetic Model Adjustment ................................................................................................ 79
5.3.1 H-abstraction from The Fuel ..................................................................................... 79
5.3.2 Fuel Radical Derived C-C and C-O Scissions........................................................... 83
5.3.3 Fuel Unimolecular Decomposition ............................................................................ 85
5.3.4 EL Concerted Elimination Reaction ......................................................................... 87
5.4 Simulation Results from Updated Kinetic Mechanism .................................................. 88
Chapter 6 CONCLUSION AND FUTURE WORK .....................................................94
REFERENCES .................................................................................................................97
SUPPLEMENTAL MATERIAL ..................................................................................109
Low-temperature Oxidation Chemistry .............................................................................. 109
Introduction to Kinetics and Thermodynamics .................................................................. 112
9

LIST OF FIGURES

Figure 2.1: Ethanol elimination reaction ...........................................................................18


Figure 2.2: (a) Transesterification from triglyceride and methanol to esters and glycerin.
R1, R2, R3 represent hydrocarbon chains of fatty acid [24]. (b) Esterification of
free fatty acid in oil with a catalyst sulfuric acid, R1 as a hydrocarbon chain. .....20
Figure 2.3: (a) methyl butanoate (b) ethyl propanoate unimolecular decomposition.
Oxygens from 1 to 2 in green, carbons from 1 to 4 in red. Ea for each transition
state was calculated by El-Nahas et al. with several techniques in Gaussian
software [29]. .........................................................................................................21
Figure 2.4: Pathways from biomass to levulinic acid via acid-catalyzed hydrolysis
process [33]. ...........................................................................................................22
Figure 3.1: JSR with crossed nozzles configuration scheme. ............................................25
Figure 3.2: Scheme of the experimental set-up at KAUST. ..............................................32
Figure 3.3: Scheme of homemade vaporizer (not to scale). ..............................................33
Figure 3.4: Scheme of microprobe for fuel introduced to JSR pre-mixed zone (not to
scale). .....................................................................................................................35
Figure 3.5: Scheme of two heating zones prior to JSR, orange region indicates preheating
zone (373-493 K); red region represents furnace heating zone (> 500 K). ...........36
Figure 3.6: Jet-Stirred reactor scheme and temperature measurement at different
locations. Temperature points separated by10 mm. SP represents set point where
reactor is monitored and controlled. ......................................................................38
Figure 3.7: Dashed lined are temperature homogeneity oven testing of jet-stirred reactor
at 650, 800, 950 and 1100K. Fig. 3.6, including position 0 as set-point spot, is
portrait positioned. .................................................................................................39
Figure 3.8: Schematic of jet-stirred reactor with molecular beam sampling system [93]. I
and II are differential pumping and ionization chambers with mass spectrometer,
respectively. 1, sVUV beamline; 2, to turbo molecular pump; 3, molecular beam
from sampling gas mixture; 4, reflectron time-of-flight mass spectrometer; 5, ion
trajectory; 6, MCP; 7, quartz cone-like nozzle; 8, jet-stirred reactor. ...................42
Figure 3.9: Syringe extraction scheme. The red sampling port is a septum used for the GC
injector. ..................................................................................................................44
Figure 3.10: GC instrument scheme, including carrier gas as helium, gas injector,
stationary-phase column, 3-way splitters to different detectors. ...........................47
Figure 3.11: GC injector scheme [23]................................................................................47
Figure 3.12: Plumbing diagram in GC-RGA. V1 and V5 are ten-port valves with
backflush function. V2, V3 are six-port valves with functions of column isolation
to TCD, column sequence reversal with backflush of the pre-column. V4 is the
10

second six-port valve with a loop for sampling upstream between V1 and V5
[97]. ........................................................................................................................49
Figure 3.13: Flowsheet of light hydrocarbon gases in valve one and two prior to
backflush process. ..................................................................................................50
Figure 3.14: Flowsheet of valve two as isolation valve is switched on, valve one is off
with backflushing. ..................................................................................................51
Figure 3.15 Flowsheet of hydrocarbons in valves three and four, before backflush .........51
Figure 3.16: Flowsheet of backflush and reverse flow as valve three is switched on. ......52
Figure 3.17: Simplified scheme of Mass Spectrometer Detector with electrical ionization
method as the ion source. .......................................................................................53
Figure 3.18: Mass spectrum of ethyl levulinate from NIST library. .................................54
Figure 3.19: A basic design of flame ionization detector [98]. .........................................55
Figure 3.20: Bridge circuit used in two-cell detector. R3 is used for sample flow detection
from column. R4 measures a reference flow from nitrogen or helium [96]. .........57
Figure 4.1: Fuel injection probe I.D. 1.6 mm with 0.7 mm exit. From top to bottom:
without sampling probe; with sampling probe; with sampling probe and pumping
device. ....................................................................................................................61
Figure 4.2: Experimental results for n-heptane oxidation with 0.5% fuel concentration, 1
atm, 2 s residence time and equivalence ratio 1. Figure 4.1-a without sampling
probe, 4.1-b with sampling probe, respectively. Configurations for blue and red
dots in portrait position. From left to right, from top to bottom: n-C7H16, O2,
CO2, and CO. .........................................................................................................64
Figure 4.3: Comparison of results of n-heptane, O2, CO, and C2H4 (top to bottom, left to
right) with different experimental configurations. .................................................65
Figure 4.4: Experimental results of cyclohexene pyrolysis in JSR. Only ethylene and 1,3-
butadiene quantified under 1 atm, 0.1% fuel and 1s residence time from 750 to
1025 K....................................................................................................................66
Figure 4.5: Experimental results of cyclohexene pyrolysis with sampling probe and
pumping device configuration with identical conditions and modeling as Fig.
4.4...........................................................................................................................66
Figure 4.6: JSR species concentration profiles (points) and simulations, using updated
mechanism from Atef et al. [102] for 0.5% fuel and 2 s residence time under
atmospheric pressure. .............................................................................................68
Figure 5.1: (a) Bond dissociation energy of ethyl propanoate (EP) (kJ/mol). Black
numbers are C-C or C-O BDE, red numbers are C-H BDE [29]. α-carbon is
designated to the carbon next to carbonyl group (C(=O)). (b) Radical stabilization
by the carbonyl group. ...........................................................................................70
Figure 5.2: Chemical structure and bond dissociation energy (BDE) of (a) ethyl levulinate
calculated by Dooley et al. [123], (b) ethyl propanoate (EP) [122] and (c) methyl
11

ethyl ketone (MEK) [121]. Red numbers represent C-H BDE, black numbers are
C-C or C-O BDE. (Unit: kcal/mol) ........................................................................71
Figure 5.3: GC-MS spectrum of EL oxidation under 850K in JSR unknown species
identified with MS detector. From A to G: A: nitrogen; B: ethanol; C: methyl
vinyl ketone; D: ethyl acrylate, E: angelica lactone, F: ethyl levulinate and G:
levulinic acid. .........................................................................................................76
Figure 5.4: Temperature profiles of ethyl levulinate (EL) oxidation, with 0.5% fuel, 1 bar,
2 s residence time and equivalence ratio from top to bottom: φ= 0.5, 1, 2............78
Figure 5.5: Radicals or fuels nomenclature of Ethyl levulinate, ethyl propanoate, and
methyl levulinate. ...................................................................................................81
Figure 5.6: Rate coefficient comparison of H-abstraction from fuel. Solid line: original
rates adopted from other molecules; dashed line: rates from the methyl levulinate
theoretical kinetic study [44]. ................................................................................82
Figure 5.7: Rate coefficient comparison of EL3J decomposing to smaller species via
beta-scission at different C-C bond positions. .......................................................85
Figure 5.8: Rate coefficient comparison of EL unimolecular decomposition, including C-
C and C-O bond scission........................................................................................86
Figure 5.9: Rate coefficient comparison in two different studies. Black line is EP
oxidation mechanism [27], used in this study. Red line is shock tube experimental
results [45] .............................................................................................................88
Figure 5.10: Updated mechanism of EL oxidation with experimental data identical to Fig
5.4. Solid lines represent updated model. Dashed lines are from previous
simulation...............................................................................................................91
Figure 5.11: Reaction path analysis for EL/O2 mixture at 1atm, 800 K (red) and 1000 K
(black), φ = 1. Percentages represent decomposition percent from parent to
daughter species. ....................................................................................................92
Figure 5.12: Sensitivity coefficients for 0.5 % ethyl levulinate in N2 under (a) 800 and
(b) 1000 K at 1 bar, φ = 1. .....................................................................................93
12

LIST OF TABLES

Table 2.1: Heating values of ethyl levulinate and other typical fuels. ...............................23
Table 3.1: Comparison of experimental conditions for jet-stirred reactor facilities in
CNRS and Nancy. ..................................................................................................26
Table 3.2: Literature review of experimental data of biofuel oxidation in JSR. ...............26
Table 3.3: Physical properties of the fuel tested and measured. ........................................34
Table 3.4: Runtime event for all valves in RGA. ..............................................................52
Table 3.5: Contributions to the effective carbon number [99]...........................................55
Table 4.1: JSR experimental parameter comparison between Nancy and KAUST ..........59
Table 5.1: Ethyl levulinate and ethyl propanoate mechanism rate coefficients for H
abstraction by OH radical; cm3 mol-1 s-1 cal-1 units. ..............................................81
Table 5.2: Rate coefficients for fuel radical derived C-C and C-O scissions, only EL1J,
EL3J and EL4J were altered. .................................................................................83
Table 5.3: Rate coefficients for EL unimolecular decomposition reactions. .....................85
13

Chapter 1 INTRODUCTION

Today’s average global temperature is the warmest in recorded history; studies cite

the cause as cumulative carbon dioxide (CO2) emissions [1]. Because petroleum-derived

fuels are the main source of CO2 emissions, alternative renewable energy options are

gaining increasing attention. As a transition from non-renewable (petroleum derived) fuels

to carbon-neutral energy, biofuels are the only feasible option for replacing conventional

transportation liquid fuels used in internal combustion engines, due to their liquid nature

portability, renewability, high combustion efficiency, and reduced net greenhouse gas

emission [2]. However, to avoid the ethical issues and social effects of first generation

biofuels, the application of second or third generation biofuels has become a more viable

option. Cellulosic biomass, a second generation biofuel, is the most abundant renewable

resource available [3]. With several pretreatments and hydrolysis conversion, ethyl

levulinate is one of the main end-products derived from lignocellulosic biomass. When

tested in diesel engines, blends of ethyl levulinate with diesel fuels showed promising

results [4]; however, the detailed combustion mechanism of this promising biofuel has not

been fully understood or developed.

The objective of this study is to develop a detailed chemical kinetic mechanism for

ethyl levulinate, one of the most promising biofuels. Dr. Stephen Dooley has completed a

preliminary model development of ethyl levulinate. Several reaction classes used analogies

to two mechanisms--ethyl propanoate and methyl ethyl ketone--since their mechanisms

were validated with other ideal reactors. To achieve validation of the developed mechanism,

an ideal reactor (the jet-stirred reactor (JSR), similar to a continuous stirred-tank reactor

(CSTR) but used for gas-phase kinetic study [5]) was implemented and constructed at the
14

Combustion and Pyrolysis Chemistry Laboratory. Before measuring the end-gas product

profiles of the target biofuel, the JSR model should be validated by conducting oxidation

and pyrolysis experiments with several reference fuels, e.g. n-heptane, iso-octane and

cyclohexene. Once the setup validation has been achieved, the oxidation of ethyl levulinate

can be studied. Hydrocarbons were detected and quantified using a gas chromatography

equipped with a flame ionization detector and a thermal conductivity detector

(GC/FID/TCD).

Further detailed adjustments have been made to the existing EL mechanism based

on comparison of experimental and simulation results. Eventually, reaction flux and

sensitivity analyses of ethyl levulinate were performed to describe the influence of several

fuel-related reactions.
15

Chapter 2 BACKGROUND RESEARCH

2.1 Introduction to Bio-derived Fuels

In December 2015 the United Nations Framework Convention on Climate Change

(UNFCCC) proposed a new international climate agreement to prevent natural catastrophes

caused by greenhouse gas (GHG) emissions; a conference of the parties (COP21) was held

in Paris, France. Their aim was worldwide reduction of GHG emissions and petroleum

dependence. An agreement was made to avoid the rise of global average temperatures

above 2 oC by the end of the 20th century. The four main ways to meet this goal included

carbon capture and storage, energy efficiency, nuclear power and renewable energy [6].

For renewable energy, biofuels, are the most viable option, produced from

biological products like biomass, liquid and gaseous fuel [7, 8]. Their main advantages are

life-cycle energy and reduced greenhouse gas emissions. Hossain et al. [9] concluded that

the total input energy required to produce one MJ of soybean oil is 0.31 MJ, while the

energy needed for one MJ of fossil fuel is 1.2 MJ. In addition, life-cycle CO2 emission is

evaluated as zero, since the plants, as the sources of biofuels, absorb CO2 during

photosynthesis. One of the most valuable applications for biofuel is in combustion engines,

where liquid phase bio-oils can be used after blending with conventional hydrocarbons.

2.2 Biofuel Use in Engines

Biofuels can be used in both spark-ignition and compression-ignition transport

engines [10]. The requirements for alternative transport fuels are that they be in liquid

phase and miscible with conventional hydrocarbons. The molecular structure of biofuels is

similar to petroleum-derived fuels, and engines require little or no alteration to

accommodate them. Molecular structure is decisive in biofuel application; they can be


16

classed either as biodiesel or bio-gasoline for use in compression-ignition and spark-

ignition engines, respectively.

In compression-ignition (CI) engines, fuel quality evaluation is based on auto-

ignition property. The cetane number (CN) can be used as an index of fuel auto-ignition

property, and the value for different fuels is derived from two primary reference fuels: n-

hexadecane (CN = 100) and 2,2,4,4,6,8,8-heptamethylnonane (iso-octane, CN = 15). The

higher the CN the higher the reactivity, as shown by short ignition delay time. Recently,

Hossain et al. [9] and Ramadhas et al. [11] tested a range of raw plants and vegetable oils

in CI engines, measuring and reviewing several physical properties of bio-oils, like cloud

point, viscosity etc., and evaluating their pollutant emissions. Karanja oil (CN = 47) was

one example tested [12]. Comprised of several fatty acids such as oleic acid

(CH3(CH2)7CH=CH(CH2)7COOH), it has a long aliphatic chain with one double bond in

the middle of its molecular structure. By blending Karanja methyl ester oil with diesel fuel,

exhaust gases were analyzed and results indicated that Karanja fuel blends reduced CO,

hydrocarbon and NOX exhaust emissions [13].

Spark-ignition (SI) engines, on the contrary, require fuels that are less reactive in

order to avoid engine knock. Pre-ignition is a phenomenon in which unburned fuel-air

mixtures spontaneously ignite before the spark timing. This unfavorable effect is followed

by engine knocking that results in reduced efficiency and even engine damage [14]. Some

biofuels, such as bio-derived ethanol, are proven to be better knock-resistant fuels than

typical gasoline fuels; and fuel additives with better knock resistance lead to higher

compression ratios and increasing the engine efficiency [14]. Therefore, mixing ethanol,

or other bio-derivatives as a typical octane booster with gasoline, appears to have an


17

important beneficial effect. The research octane number (RON) represents the tendency of

fuels to resist auto-ignition; in ethanol, the RON is defined as 109 [15]. The primary

reference fuels used for characterizing the RON scale are n-heptane (RON = 0) and 2,2,4-

trimethyl-pentane (iso-octane, RON = 100). Wei-Dong et al. [16] investigated the engine

performance and emissions of different mixtures of ethanol and gasoline (0%-30% ethanol);

their experimental results showed that a higher blending percentage significantly decreased

the emission of CO and unburned hydrocarbon, but increased CO2 emission quantity; this

is similar to emissions from biodiesel-diesel fuel blends. However, in terms of heat release

rate, calorific value and pressure in the combustion chamber, the engine performance of

gasoline-ethanol blends does not achieve the same level as fossil fuels [17].

2.3 Bio-derived Fuel Combustion Chemistry

Combustion is a series of species production and consumption that includes

thousands of reactions, eventually releasing heat in the form of power for engine use. These

reaction classes include decompositions and re-combinations of both pyrolysis and

oxidation. Because they contain oxygen atoms that promote engine oxidation and achieve

nearly complete combustion, biofuels are treated as partially oxidized hydrocarbons, even

to their molecular structure. However, disparities in functionality produce unpredictable

effects on basic combustion chemistry, directly impacting fuel blending and engine

selection. In the following section, another classification of biofuels is discussed which

includes alcohol and carboxyl in its functional group.

2.3.1 Alcohols Chemistry

Among all the alcohols extracted from biomass, ethanol has been the most widely

used. Ethanol is primarily derived from corn and sugar through a fermentation process; in
18

2015 global ethanol production surpassed 20 billion gallons [18]. The United States and

Brazil are the two major ethanol fuel producing countries, accounting for 85% of ethanol

worldwide. Yet social and ethical issues, such as stable food pricing in developing

countries, dictate that edible crops be grown to produce second or third generation biomass

for the energy section [19] rather than being converted to ethanol.

Regarding its molecular structure, some unimolecular decompositions are critical

to high-temperature combustion. With hydroxyl functionality, the oxygen atom in alcohol

attracts a hydrogen atom bonded to a carbon (Fig. 2.1), forming hydrogen bonding that

causes direct molecular elimination reactions (because hydrogen bonding is strong enough

to break the C-H and C-O bonds), and producing alkene and water. However, because of

the electronegativity that promotes hydrogen abstraction from other radicals during

combustion, the oxygen atom weakens neighboring C-H bonds.

ethanol ethylene + water

Figure 2.1: Ethanol elimination reaction


2.3.2 Biodiesel Chemistry

Biodiesels are mostly extracted from vegetable oils and animal fats. Typically,

mono-alkyl esters are produced after transesterification from long-chain fatty acid and

alcohols; the carbon number ranges from 12 to 18. However, because of their high viscosity

and combustion-relative problems [20], neat vegetable oils are not appropriate for direct

feeding into diesel engines. Farwood et al. evaluated short- and long-term effects, such as
19

plugging and gumming of filters, lines and injectors, and coking of a piston injector due to

high viscosity [21]. One way to improve the fuel properties of vegetable oils and avoid fuel

polymerization is to produce monoesters from triglyceride and methanol using a base as a

catalyst (Fig. 1.2 (a)), also known as transesterification. These monoesters usually contain

long aliphatic chains of 12 to 18 carbons with 10 to 11% oxygen weight percentage [22],

and they exhibit high CN for diesel engine use.

Free fatty acids (FFA) are also derived from biomass such as oleic acid in Karanja

oil. FFAs easily react with the trans-esterifying catalysts by forming soap and water,

complicating the conversion process and reducing the yield of esters. Pretreatment is

required if the FFA in oil is higher than 5%. To solve the problem, FFAs can undergo

esterification with methanol and a strong acid (such as sulfuric acid), so that the hydroxyl

group is replaced by an alkyl group and converted to methyl ester and water, described in

Fig. 1.2 (b) [23].

(a) O
O
=
=

CH3 – O – C –R1
CH2 – O – C –R1 CH2 – OH

O O
=

CH – O – C –R2 + 3 CH3OH CH3 – O – C –R2 + CH – OH

O
O
=

CH2 – O – C –R3 CH2 – OH


CH3 – O – C –R3

(b) O H2SO4 O
+ CH3OH + H2O
=
=

HO – C –R1 CH3 - O – C –R1


20

Figure 2.2: (a) Transesterification from triglyceride and methanol to esters and glycerin.

R1, R2, R3 represent hydrocarbon chains of fatty acid [24]. (b) Esterification of free fatty

acid in oil with a catalyst sulfuric acid, R1 as a hydrocarbon chain.

Some short carbon-chain esters can be blended with gasoline as octane boosters [25].

At the molecular level, a carboxyl group is the significant difference between some biofuels

and conventional gasoline or diesel fuel. Under high temperature combustion conditions

(> 700 K), unimolecular decomposition is critical and its rate coefficient must be precisely

determined. In general, high temperatures help alkyl esters to decompose to carboxylic

acids and olefins through four or six-member transition state rings [26]. Taking methyl

butanoate (MB) and ethyl propanoate (EP) for example, 50% of MB fuel is consumed at

10 bar, 900 K, through a six-centered concerted elimination reaction, shown in Fig. 2.3 (a).

The most favorable hydrogen site to be attacked by the carbonyl group is on C4, due to the

six-membered transition state ring. When the C4-H bond breaks, a beta-scission reaction

occurs via C2-C3 bond cleavage which produces ethylene and methyl ethanoate. EP has a

similar reaction (Fig. 2.3 (b)). The O1 connected to C1 tends to bond with the hydrogen on

C3, forming another six-centered transition state. Then the C3-H and O2-C2 bonds break

followed by O1-H single and C2=O2 double bonds formation. Comparing the two

transition states, the activation energy of the six-centered ring for EP is 75.53 kJ/mole

lower than MB. This decrease in activation energy results in higher reactivity and faster

ignition of the EP molecule [27-29].


21

(a) 1 4 1

2 3 Ea = 209 kJ
3 1 2
4 2 1
2
1
4 3
1

2 2

(b) 1 3

1 2
2
1
1 3
2 3
2 Ea = 284.5 kJ
1 1
2 2

Figure 2.3: (a) methyl butanoate (b) ethyl propanoate unimolecular decomposition.

Oxygens from 1 to 2 in green, carbons from 1 to 4 in red. Ea for each transition state was

calculated by El-Nahas et al. with several techniques in Gaussian software [29].

2.4 From Lignocellulosic Biomass to Ethyl Levulinate

Aside from chemistry moieties, biofuels can be divided into several categories, based

on their feedstock. The Renewable Fuel Standard (RFS) program [30], assigned to replace

the current petroleum fuel from U.S. national policy, categorizes all biofuels into four

sections: Biomass-based diesel, cellulosic biofuel, advanced biofuel and total renewable

fuel. Cellulosic biofuels, such as wood or straw, achieve lifecycle GHG emission

reductions as low as 60%, the highest percentage among the other three sources; it is
22

considered the most promising alternative natural resource for the sustainable supply of

fuels and chemicals in the future. The amount of cellulose production is estimated at

approximately 2 × 109 tons per year [31]; and these products can be utilized in food, fiber,

and pharmaceutical applications. With acid-catalyzed dehydration, cellulose is

decomposed into smaller molecules: glucose (or hexose), by breaking the C-O bonds

connected to each cellulose unit. Furthermore, the hydrolysis of glucose produces 5-

hydromethylfurfural (5-HMF), [32]. Finally, heat and an acid catalyst convert 5-HMF to

levulinic acid (LA) as the main product and formic acid as a byproduct. The pathways from

feedstock to the desired product--levulinic acid--are described in Fig. 2.4.

Hemicellulose
Glucose
Biomass Cellulose

Lignin -3H2O +H+

+2H2O

+H+, -HCOOH

Levulinic acid 5-hydroxymethylfurfural (HMF)

Figure 2.4: Pathways from biomass to levulinic acid via acid-catalyzed hydrolysis

process [33].

In 2000, Bozell [3] proposed that LA has many potential applications for the food

industry, electronics, drug delivery and fuel, and in plasticizers [34]. In the energy sector,

the most promising diesel and gasoline additives from cellulosic feedstock are the esters of

LA [35] or 2-methyltetrahydrofuran (2-MTHF) [36], whose flame combustion mechanism

is well understood [37]. Using an acid catalyst, levulinic esters can be synthesized via the
23

esterification of LA with alcohols. Ethyl levulinate (EL), is one of these esters, and it is

produced predominantly over methyl and propyl esters, since ethanol is widely produced

and commercialized.

EL is considered a potential fuel surrogate after blending with diesel fuels for several

reasons: First, the physical properties of EL are similar to that of biodiesel fatty acid methyl

esters (FAME). Joshi et al. [38] mentioned that blends of EL and cottonseed oil methyl

esters, or poultry fat methyl esters, could improve cold flow properties by decreasing cloud

and pour point, beneficial for transportation in high latitudes. Second, Christensen and

coworkers [4] reported that EL may be used as a fuel lubricant to expand the lifetime of

engine components. Klemm et al. [39] also suggested that a diesel engine could continue

to operate, without modification to the engine components, when the percentage of EL in

the blend was 20%. Moreover, the lower heating value (LHV) of EL was investigated and

compared with other fuels (see Table 2.1), where LHV is defined as the heat released from

the fuel combustion (initially at 25 °C), and the temperature of the combustion products is

returned to 150 °C. Note that the latent heat of vaporization of water in the reaction

products is not considered [40]. It is worth noting that blending 3% of EL with diesel fuel

caused 0.5 MJ/kg heating value loss, but reduced CO and particulate matter emission

significantly. Similarly, blends of butyl levulinate with diesel fuel can decrease soot

emissions by as much as 95% [41]. However, Christensen questioned the feasibility of EL,

because of its low CN and poor solubility in diesel blends at low temperatures.

Table 2.1: Heating values of ethyl levulinate and other typical fuels.

Fuel Lower Heating value (MJ/kg) Ref.

Ethyl levulinate 24.8 [42]


24

Butyl levulinate 27.4 [4]

Methanol 20.1 [43]

Ethanol 26.9 [43]

Butanol 34.4 [43]

Diesel fuel 43.2 [42]

Biodiesel 40.0 [42]

Conventional gasoline 43.4 [43]

Natural gas 47.1 [43]

3% EL in Diesel 42.7 [42]

The only theoretical and experimental studies concerning levulinic esters have been

conducted by Thion et al. [44] and AlAbbad et al. [45]. Thion carried out the calculation

of methyl levulinate (ML) oxidation by OH and CH3 radicals, forming ML radicals and

exploring the possible pathways from ML to smaller species such as methyl vinyl ketone

and methyl acrylate. AlAbbad conducted the measurement of EL unimolecular

decomposition reaction to LA and ethylene under different pressure conditions at high

temperature (900 to 1500 K). (Some rate coefficients in this study were adopted from their

results and will be further mentioned in chapter 5.)


25

Chapter 3 JET-STIRRED REACTOR VALIDATION

Experimental work focused exclusively on constructing a new jet-stirred reactor

(JSR) facility at King Abdullah University of Science and Technology (KAUST), Clean

Combustion Research Center (CCRC).

Figure 3.1: JSR with crossed nozzles configuration scheme.

3.1 Literature Review on Biofuels Combustion in JSR

The JSR was first introduced by Villermaux [5]; it is suitable for gas-phase kinetic

studies because it is taken as a zero-dimension reactor, since fuel transporting properties

need not be considered, unlike co-flow or counter-flow flame. Currently, two groups are

still validating gas-phase kinetic mechanisms with JSR facilities in France. In Dagaut’s

model (CNRS France), the volume of the reactor is 35 cm3, equipped with four nozzles, 2

mm in diameter, to achieve perfect stirring. Liquid fuel is injected at constant flow rate,

using a HPLC pump connected to a vaporizer. To reach pressures of 40 bar, the inlet and

outlet pipelines are inside a stainless steel jacket, and to prevent temperature gradients, the

setup is insulated with heating tapes. Nitrogen gas is the carrier gas. Species are analyzed

with several gas chromatography machines installed with different columns, flame
26

ionization detectors and thermal conductivity detectors [46-48]. Another JSR setup,

established in Nancy, uses a reactor volume of 95 cm3. The number of preheating zones in

this setup is divided into two for progressive heating of the mixture gas [49]. Table 3.1,

presents experimental setup comparisons of JSR facilities in different research institutes.

Table 3.1: Comparison of experimental conditions for jet-stirred reactor facilities in

CNRS and Nancy.

JSR in CNRS# JSR in Nancy*


Pressure (bar) 10-40 1.067
Residence time (s) 0.01-3 2
Equivalence ratio φ 0.15-4.0 0.5, 1.0, 2.0, 4.0
Temperature range (K) 900-1200 500-1000
Fuel concentration (%) 0.15-0.5 0.5
Reactor volume (cm3) 35 95
Number of preheating zones 1 2
Carrier gas N2 He
Sampling method online and offline online and offline
# parameters were referred to [46].
* parameters in this table were adopted from n-heptane oxidation measurement [50].

A summary of available studies of fundamental reaction kinetics regarding all biofuels,

including alcohols, ethers, ketone, and methyl/ethyl esters is provided in Table 3.2. The

number of ethanol oxidation studies far exceeds that of other biofuels.

Table 3.2: Literature review of experimental data of biofuel oxidation in JSR.

Temperature Pressure
Fuel Equivalence ratio Reference
range (K) (bar)
27

Singh et al. 1979


1870 0.93 0.7-1.4
[51]
Methanol
Burke et al. 2016
800-1200 10 0.2-2
[52]

Togbe et al. 2009


Methanol, M85 surrogate 770-1140 10 0.35-2
[53]

Dagaut et al.
Ethanol ~1000 1 0.2-2
1992 [54]

Dagaut and
Ethanol, E85 surrogate 800-1100 10 0.6-2
Togbe, 2008 [55]

Galmiche et al.
n-propanol 770-1190 10 0.35-2
2011 [56]

Dagaut et al.
750-1100 10 0.5-2
2009 [57]
n-butanol
Sarathy et al.
850-1250 1 0.25-2
2009 [58]

n-butanol/ n-heptane Dagaut and


530-1070 10 0.5 and 1
mixture Togbe, 2009 [59]

Togbe et al. 2010


2- and iso-butanol 770-1250 10 0.275-4
[60]

Serinyel et al.
2-methyl-1-butanol 700-1200 10 0.5-4
2014 [61]
28

423, 600- Togbe et al. 2010


n-hexanol 10 0.5-3.5
1100 [62]

Dayma et al.
iso-pentanol 530-1220 10 0.35-4
2011 [63]

423, 770- Togbe et al. 2011


n-pentanol 10 0.7-1.4
1220 [64]

Mze-Ahmed et al.
n-hexanol/ jet A-1 mixture 560-1030 10 0.5-2
2012 [65]

Cai et al. 2015


n-octanol 500-1200 10 0.5-2
[66]

Gaïl et al. 2007


800-1350 1 1.13
[67]

Sarathy et al.
850-1350 1 1
2007 [68]
Methyl butanoate
Gaïl et al. 2008
850-1400 1 0.375, 0.75
[69]

Hakka et al. 2010


800-850 1 0.5-1
[70]

Gaïl et al. 2008


Methyl crotonate 850-1400 1 0.375-1
[69]

Dayma et al.
Methyl hexanoate 500-1100 10 0.5-1.5
2008 [71]
29

Dayma et al.
Methyl heptanoate 550-1150 10 1-2
2009 [72]

Dayma et al.
Methyl octanoate 800-1350 1 0.5-2
2011 [73]

Glaude et al.
Methyl decanoate 500-1100 1.06 1
2010 [74]

Hakka et al. 2009


Methyl palmitate 500-1000 1 1
[75]

Bax et al., 2010


Methyl oleate 550-1100 1 1
[76]

Rapeseed methyl esters Dagaut et al.


800-1400 1-10 0.25-1.5
(RME) 2007 [77]

Dagaut et al.
RME/kerosene mixture 740-1200 1 0.5-1.5
2007 [78]

Metcalfe et al.
Ethyl propanoate 750-1100 1 0.3-1
2009 [27]

Dayma et al.
Ethyl pentanoate 560-1160 10 0.6-2
2012 [79]

Commercial B30 fuel and Ramirez et al.


560-1030 6 and 10 0.25-1.5
B30 surrogate fuel 2011 [80]

Serinyel et al.
n-methyl butanal 500-1200 10 0.35-4
2017 [81]
30

Rodriguez et al.
n-hexanal 475-1100 1 0.25-2
2017 [82]

Thion et al. 2016


Cyclopentanone 730-1280 1, 10 0.5-2
[83]

Thion et al. 2017


Methyl ethyl ketone 700-1150 10 0.5-2
[84]

Shahla et al. 2017


Diethyl carbonate 680-1220 1 0.5-2
[85]

Curran et al. 1998


Dimethyl ether 800-1300 1, 10 0.2-2.5
[86]

Daly et al. 2001


Dimethoxymethane 800-1200 5.07 0.444-1.778
[87]

Methyl/ethyl tertiary butyl Glaude et al.


750-1150 10 0.5-2
ether 2000 [88]

3.2 Experimental Set-Up at KAUST

The JSR-type selection follows rules established by Ayass et al. [89], who

recommended that as the experiment with larger residence times (from 0.5 to 5 seconds),

the JSR should be selected with a larger diameter and crossed nozzles configuration for

ensuring better mixing; for this reason, a reactor volume of 76 cm3 is included in this model

(see Fig. 3.1). The nozzle diameter is 0.33 mm.

The setup was validated using several primary reference fuels such as n-heptane, which

has been studied using a setup and experimental conditions similar to those used in Nancy.
31

The iso-octane was also tested as representative of a high-octane-number species.

Afterward, the setup was used for conducting EL oxidation and compared with simulation

results.

The JSR in this study was made from quartz glass. The reactor was connected to

both upstream and downstream quartz tubes of diameter 25.4 mm. The values of residence

time for all experiments in this study were set to two seconds and operated under

atmospheric pressure. The temperature of the electrical furnace can reach 1500 K

maximum. Figure 3.2 shows a schematic diagram of the complete JSR setup. The apparatus

used includes a syringe pump injecting liquid fuel, a multi-gas controller from MKS,

coupled with three mass flow controllers, four K-type thermocouples, a homemade

vaporizer, a microprobe, an electrical heating furnace, three electrically insulated resistors,

four thermometers, and a sampling tube with a glass finger. The mixture was analyzed

using a gas chromatography (GC), coupled with a mass spectrometry detector (MSD),

flame ionization detector (FID) and a thermal conductivity detector (TCD), or a GC

designed as a refinery gas analyzer (RGA), one FID and two TCDs. Different columns,

such as DB-624, DB-1, GAS-Pro, Molecular Sieve 5A were implemented for species

separation and detection.


32

Multi-gas controller

Pressure
gauge MFC
O2 supply

N2 supply JSR Sampling tube

Pressure valve GC

Vaporizer
Thermocouple

Fuel Temperature
controller

Thermometer
Syringe pump Electrical heating furnace

Figure 3.2: Scheme of the experimental set-up at KAUST.


33

3.2.1 Fuel Injection System

The fuel was stored in a 25-milliliter syringe provided by SGE Analytical Science

and injected using a syringe pump from New Era Pump Systems Inc.; this syringe pump

can be operated with a wide volume flow rate, ranging from μl per hour to ml per minute,

with syringes of different volumes. Fuel in liquid phase was injected into a homemade

vaporizer, which was heated to a specific temperature point 30 K higher than the boiling

point of the fuel in use (Table 3.3). Fig. 3.3 shows the gap between the fuel injection pipe

and the chamber, filled with very small stainless-steel chips for better heat transfer. The

temperature was measured using a stainless-steel K-type thermocouple inserted between

the chamber jacket and the heating tapes.

vaporized fuel
diluted in N2

heating
tape

stainless N2 as
steel chips carrier gas

liquid fuel
in I.D: 1.6 mm pipeline

Figure 3.3: Scheme of homemade vaporizer (not to scale).


34

Table 3.3: Physical properties of the fuel tested and measured.

Vaporizer setpoint Vapor density


Fuel tested Boiling point (°C)
(°C) (kg/m3)
Ethyl levulinate 206-207 240 5.948
iso-octane 99.3 130 3.930
n-heptane 98.4 130 3.070

Nitrogen (N2) was used as a carrier gas, transporting the fuel to the microprobe (see

Fig. 3.4 for microprobe dimensions). At the microprobe, the fuel was mixed with O2 10

mm prior to the JSR nozzle entrance. After the desired residence time in the JSR, the

unreacted fuel/end-products flowed to a detection device. All transfer lines were heated by

insulated resistors coupled with thermocouples and monitored using temperature

controllers to prevent condensation from the outlet of vaporizer to the microprobe

(expressed as preheating zone in Fig. 3.5). The temperature was set 20 K below the

vaporizer temperature. As the mixture traveled into the furnace heating zone, it flowed

through a 70 mm-long microprobe, to avoid temperature gradients, and then entered the

JSR. The residence time of the fuel in the microprobe was calculated; the speed of the gas

mixture could reach high velocity over 400 cm/s at 700 K because the capillary had a 0.8

mm inner diameter which resulted in 0.0175 s residence time of the fuel traveling through

the tube. Since this is a small value it can be assumed that it did not significantly affect the

results. These parameters were calculated based on Eq. 3.1 and 3.2. The volume of

microprobe is expressed as 𝑉; r is the radius; 𝑙 is the length at the same temperature as

reactor; 𝑄 is the total volume flow rate of fuel/N2 mixture and 𝜏 is the residence time.
35

𝑉 = 𝜋𝑟 2 𝑙 (Eq. 3-1)

𝜏 = 𝑉/𝑄 (Eq. 3-2)

Oxygen was introduced outside the microprobe and diluted with N2. The two flows met 10

mm away from the nozzle entrance, defined as the gas pre-mixing zone. The time that the

fuel/oxygen/nitrogen mixture remained in the pre-mixing zone was estimated by Eq. 3.1

and 3.2 and the calculation was near 0.05 s at 700 K. In total, the residence time of the fuel

in the furnace zone (microprobe + pre-mixing zone) prior to the JSR was approximately

3.5 % of the residence time in the reactor. The conclusion can be made that the residence

time of the fuel in both microprobe and the premixing zone with oxidizer was relatively

insignificant (3.5%) in comparison with the residence time 2 s of the reactor.

Fuel and O2 were premixed at the inlet of the fused silica reactor (pre-mixing zone),

which achieved good fuel/O2 premixing. Although this resulted in the mixture starting to

react before entering the JSR, it was preferable to poor fuel/O2 mixing. Several high-

temperature-resistance O-rings, with sizes ranging from a quarter to one inch were used for

the adaptors between the glass and metal tubes.

O.D: 2 mm O.D: 6.35 mm


I.D: 0.8 mm I.D: 2 mm

320 mm

Figure 3.4: Scheme of microprobe for fuel introduced to JSR pre-mixed zone (not to

scale).
36

70 mm 10 mm

Fuel/N2
O2/N2

Preheating zone Furnace heating zone

Figure 3.5: Scheme of two heating zones prior to JSR, orange region indicates preheating

zone (373-493 K); red region represents furnace heating zone (> 500 K).

3.3 Verification of Criteria for KAUST JSR

Lignola [90] designed another jet-stirred reactor in Italy and established some

criteria to ensure that the gas mixture in the jet-stirred reactor reached homogeneity; the

design was thoroughly re-examined by Herbinet and Dayma [49] from Nancy. Along with

the guidelines, three criteria are listed below:

1. Jets from the four nozzles must be turbulent.

2. The four jets must provide good mixing of the entire gas phase in the reactor; the jets

must provide an intense internal recycle system.

3. The velocity of the jets at the outlet of the nozzles must not exceed the speed of sound,

which could interfere with the mixing process in case of shock wave formation.

For verification, several parameters were limited to ensure that the JSR achieved

the criteria. Some of these equations are referred from Matras 1975 [91]. To meet the first

criterion, eq. 3.3 should be satisfied. 𝜏 represents residence time; η is dynamic viscosity; 𝜌

is the specific weight of the gas; 𝑅 is the radius of the reactor; 𝑑 is the diameter of the
37

nozzle and 𝐴 is a dimensional constant as a function of temperature, 0.3 in the case of air

at 293.15 K and atmospheric pressure. In this work, studying nitrogen gas at temperatures

of 700 and 1000 K: 𝜌 is 0.483 and 0.337 kg/m3; 𝑅 is 2.915 x 10-2 m; η is around 0.032 and

0.0415 mPa/s; 𝑑 is 0.3 mm and 𝐴 is π/4. Ultimately the right side of eq. 3.3 is equal to 4.26

and 2.3 s, larger than the assigned residence time 2 s in this study.

𝜌𝐴𝑅 3
𝜏≤ (Eq. 3-3)
230η𝑑

To meet the second criterion, the required ratio between radius of the reactor and

the internal diameter of the nozzles must satisfy equation 3.4. The calculated value is equal

to 97.2, allowing for perfect mixing.


𝑅
≥ 64 (Eq. 3-4)
𝑑

Of equal importance, the velocity u0 can be expressed as a function of the radius of

the reactor 𝑅, the nozzle diameter 𝑑 and the residence time 𝜏, leading to the requirement,

as shown in eq. 3-6. Gas flowing through the reactor is assumed to be nitrogen and 𝑐𝑠𝑜𝑢𝑛𝑑 ,

which can be calculated from the equation 3-5, with 𝛾 as heat capacity ratio, R as ideal gas

constant, 𝑇 as temperature and 𝑀 as molecular weight. Considering 𝑇 to be 700 and 1000

K, 𝑐𝑠𝑜𝑢𝑛𝑑 is calculated as 539.4, 644.7 m/s. Hence the minimum residence time is 0.68 and

0.57 s.

𝛾𝑹𝑇
𝑐𝑠𝑜𝑢𝑛𝑑 = √ (Eq. 3-5)
𝑀

4 𝑅3
𝜏≥ (Eq. 3-6)
3 𝑑 2 𝑐𝑠𝑜𝑢𝑛𝑑 (𝑇,𝑃)

In conclusion, experimental conditions in this case satisfied the stated criteria. The

residence time (2 s in this study), is within 0.68 and 4.26 s at 700 K and between 0.57 and
38

2.3 s at 1000 K. Also, turbulent jets, the desired recycling rate, and a flow less than that of

the sonic limit were all achieved.

3.4 Temperature Homogeneity Measurement

Temperature homogeneity is considered to be one of the most critical factors in

reaction gas-phase kinetic study, since rate coefficients are very sensitive to temperature.

In the JSR set in this study, a total length of 200 mm, starting from the upstream microprobe

to the downstream sampling probe, was placed and partially heated in the furnace.

Temperature distribution in the reactor and the tubes were measured to prove thermal

homogeneity. Different locations were measured, as shown in Fig. 3.6. Set point (SP) was

defined as the position of the ideal reactor temperature and N2 was set at a constant flow

rate during the measurements.

-40 -30 -20 -10


SP 10 20 30

Jet-stirred reactor

Electric furnace

Figure 3.6: Jet-Stirred reactor scheme and temperature measurement at different

locations. Temperature points separated by10 mm. SP represents set point where reactor

is monitored and controlled.


39

The temperature profile is presented in Fig. 3.7. Each location in the figure was

measured with 1% error and repeated more than five times. The largest deviation is located

at 650 K, which is 10 K lower than the SP. Downstream locations showed fewer

temperature gradients than upstream. Overall, the furnace maintained good temperature

homogeneity, especially under high-temperature set points.

1200

1050
Temperature (K)

900

750

600

-30 -20 -10 0 10 20 30


Position (mm)

Figure 3.7: Dashed lined are temperature homogeneity oven testing of jet-stirred reactor

at 650, 800, 950 and 1100 K. Fig. 3.6, including position 0 as set-point spot, is portrait

positioned.

3.5 Sampling Methods

Sampling is critical to the accuracy of gas-phase composition analysis. In Dagaut’s

group [49], online and offline analysis were applied for different end-product collections,

based on their physical properties. Another sampling technology--molecular beam

sampling--is effective for unstable species, especially oxygenated compounds with

relatively short lifetimes.


40

1. Online sampling: The online analysis is the most effective and useful method for

keeping the composition nearly identical to the outlet of the reactor. As species are

transported in the downstream tube connected to the JSR, a sampling sonic probe

is installed, fed into the reactor and connected to the transfer line. All tubes are

heated to prevent condensation. Through a 6- or 10-way injection valve system on

a gas chromatograph, a constant volume of sampling gas is introduced from the

loop to a column for species separation and analysis.

2. Offline sampling: Offline sampling was used to collect low vapor pressure species

in a glass vessel under very low temperature or low pressure conditions. Liquid

nitrogen is usually used as a bath matrix for cooling to the necessary low

temperatures. Condensed samples are stored in solid phase and then re-heated with

the addition of a solvent like acetone. However, the vessel can also be connected to

a differential pump. When low pressure conditions are reached in the vessel, species

with low vapor pressure can be stored in gas phase. This method was implemented

for n-dodecane pyrolysis [92].

3. Molecular beam mass spectrometry: The sampling methods mentioned above are

used mainly for stable species analysis. It has been proven that the unstable species,

such as hydroperoxide, decomposes, or reacts to other species instantly, prior to

sampling and analysis. Molecular beam sampling was first introduced by Battin-

Leclerc and Herbinet [93, 94] at the National Synchrotron Radiation Laboratory

(NSRL) in Hefei, China. (Experimental model described in Fig. 3.8.) This setup

was designed for unstable species which were analyzed using tunable synchrotron

vacuum ultraviolet (sVUV) photoionization mass spectrometry. With a quartz cone


41

and an orifice attached to a JSR, all species form a molecular beam as the pressure

drops from 105 to 0.067 Pa and are then introduced at high velocity into a vacuum

chamber under a pressure of 8.3x10-4 Pa. Horizontally a sVUV light source with

7.8 to 24 e.V tunable photon energy was introduced to ionize the sampled gases.

Eventually, in the photoionization chamber, species were accelerated into the mass

spectrometer, due to the electric field, and separated by a mass reflectron

technology. Ions were amplified with a microchannel plate (MCP) and signals were

detected with a photo-multiplier tube (PMT). A recent contribution of this

apparatus was the discovery of several third-O2-addition products under low-

temperature condition of alkane oxidation [95].


42

Figure 3.8: Schematic of jet-stirred reactor with molecular beam sampling system [93]. I

and II are differential pumping and ionization chambers with mass spectrometer,

respectively. 1, sVUV beamline; 2, to turbo molecular pump; 3, molecular beam from

sampling gas mixture; 4, reflectron time-of-flight mass spectrometer; 5, ion trajectory; 6,

MCP; 7, quartz cone-like nozzle; 8, jet-stirred reactor.

3.5.1 Downstream Sampling Design

Although the JSR is considered to be a perfect stirred reactor for gas-phase mixtures,

concentration and velocity distributions still exist in the reactor [89]. This discrepancy can

be neglected once the species concentrations are averaged after sampling quantitatively

and qualitatively. According to Herbinet’s research [49], online analysis with a sampling
43

probe installation can maintain the composition of stable species in the gas phase--even

away from the reaction zone--with the incorporation of an additional heating zone. Since

sampling is critical to the accuracy of product analysis, two methods were tested in this

study--with and without sampling tubes. The species are then transported to the detection

apparatus via either online analysis or syringe injection. All combinations are listed and

explained below:

1. With or without sampling probe: The sampling probe was composed of relatively

fragile fused silica (Fig. 3.8); it was designed for temperature measurement and

sample collection. As noted in Chapter 3.3, the thermocouple is inserted into a fused

silica glass finger next to the sampling probe at the exit of the JSR. The sampling

tube was designed with 2 mm I.D., the opposite end was designed to fit the

connection to the other stainless steel tube, similar to the sample injection

microprobe configuration. The method without a sampling probe was tested for

comparison when the sampling probe was not available for use.

2. Pumping system:

A vacuum diaphragm pump from KAMOER® was installed at the downstream port

out of the GC system for shorter retention time of the products. The volume flow

rate through the sampling probe was improved from 10 standard cubic centimeter

per minute (sccm) to over 60 sccm, causing a sextuple decrease of residence time

in the sampling probe.

3. Online analysis and gas-tight syringe injection:

Online sampling connects the downstream probe to the analytical detectors through

heated pipelines, preventing sample condensation. However, when the online


44

analysis from reactor to detectors was unavailable, the gas-tight syringe method

was implemented, depending on the configuration of the gas chromatography

system. During species transportation in the pipeline, a GC-suitable septum was

installed and locked to maintain downstream pressure and prevent leakage. The

instrumental setup is shown in Fig. 3.9. By extracting a constant volume of the gas-

phase sample (500 μl from the downstream side, with a gas-tight syringe), the

sample was injected immediately and manually through the septum at the injector

part of the GC system. Hexane and acetone were used to keep the syringe clean and

uncontaminated.

Thermocouple

Sampling port
with a septum

End-gas exhaustion

Figure 3.9: Syringe extraction scheme. The red sampling port is a septum used for the GC

injector.
45

3.6 Analytical Techniques

Two analytical apparatus were used for downstream species qualification and

quantification:

1. Gas chromatography (GC), coupled with a mass spectrometry detector (MSD),

flame ionization detector (FID) and thermal conductivity detector (TCD).

2. Gas chromatography (GC), with refinery gas analysis (RGA) valve system, a

FID, and two TCDs.

The samples were analyzed with gas chromatography using an Agilent 7890 (Agilent

Technologies, Santa Clara, CA.), equipped with a split/splitless injector. Injector operating

conditions were as follows: The injection volume was 500 μl, inlet temperature was

maintained at 280 °C to prevent condensation. Helium gas flow was used as the carrier

gas. The programmable oven was set at 50 °C for 2 min, then 14 °C/min to 195 °C for

0.6 min, then 10 °C/min to 250 °C for 5.5 min. The total run time was 23.96 min. The

analytical column used was a DB-624 capillary column (6% Cyanopropylphenyl

94% dimethylpolysiloxane; 60-m × 250-μm i.d. and 1.4-μm film thickness). The system

worked at a constant flow. The exit end of the analytical column was connected to one of

the four ports on the three-way splitter. This device divides the effluent from the analytical

column between three different detectors: (1) to the interfaced mass selective detector

(Agilent 5975 MSD), (2) to the thermal conductivity detector (TCD), and (3) to the flame

ionization detector (FID). Three deactivated fused silica capillary restrictors, located from

the three-way splitter to each of the three detectors, were installed with different lengths

and diameters: MSD restrictor (1.44-m × 0.18-mm i.d.), TCD restrictor (1.06-m × 0.18-mm

i.d.), and FID restrictor (1.06 m × 0.18 mm). The split ratio applied was 2:1:1 (MSD: TCD:
46

FID). The splitter uses an auxiliary pneumatic control module (PCM) working to keep the

pressure constant at 0.262 bar, thus maintaining the split ratio fixed in each run.

The MSD worked in PEI mode, operating in full-scan mode; monitoring was from

4 to 200 m/z. All the gases used in this work had a high purity of 99.9995% or better. The

instrument control, data acquisition, and processing were performed using Agilent MSD

ChemStation software (E.02.00.3 version).

For product identification purposes, including hydrocarbons, permanent gases (e.g.

carbon dioxide, carbon monoxide, oxygen, hydrogen, etc.), MSD was used. TCD was used

for non-organic species analysis, while the FID detector performed quantitative analyses

of hydrocarbons.

3.6.1 Gas Chromatography

The gas chromatography (GC) (Fig. 3.10) is considered to be one of the most useful

instruments for separating and quantifying volatile organic compounds in mixtures by

means of unique retention times, based on their physical properties. However, it cannot

identify molecular structures unless the retention time for the species has already been

detected using a standard solution/cylinder under the same temperature program. After

injection (Fig. 3.11), the mixture is diluted and carried by mobile phase (helium) and

introduced to a packed, or a capillary, column. Each compound acts as a sorbate; its flowing

speed and separation process are determined by the interaction of its physical or chemical

properties with the solid packing coated in the column. Complete separation can be

achieved when the column is long enough, or has the appropriate diameter. A longer

column is necessary if the peak is too broad, while a smaller diameter column provides
47

better high resolution and retention. The run time is lengthy, but this allows the peaks to

be easily separated and analyzed.

3-way Splitters

Injector Detectors

GC/MS/FID/TCD
system
Carrier gas Column
(Helium)
Oven

Figure 3.10: GC instrument scheme, including carrier gas as helium, gas injector,

stationary-phase column, 3-way splitters to different detectors.

Figure 3.11: GC injector scheme [23].


48

3.6.2 Column Selection

Generally, there are two types of column for the GC system, the packed and the

capillary columns, with different functionalities. The packed column is composed of three

parts: tube, packing retainers and packing materials, serving as stationary phases [96].

Packing materials are designed with enough surface space for compounds to adsorb, and

GC efficiency is highly relative to the polarity and selectivity of the mixture. For the

columns employed in this study, DB-624 was used to separate esters, or some oxygenated

hydrocarbons classified as having medium polarity; DB-5-MS, for separating

hydrocarbons like n-heptane, has low polarity. However, capillary columns (also referred

to as open tubular columns), separate very light and highly volatile compounds such as

methane, ethane, carbon monoxide etc., in a relatively shorter time of analysis. The

stationary phase is usually constructed using either metal or fused silica and the most

common column type is the porous-layer open-tubular (PLOT), which uses metal oxide as

the stationary phase (e.g. Al2O3-PLOT, KCl-PLOT). GS-GASPro is another useful

capillary column used in this study to separate C1-C3 hydrocarbons. A molecular sieve can

be used to separate and quantify CO, CH4, O2 and N2.

The inner diameter and length of the column are critical factors for better separation

and quantification, but optimizing one of these factors requires a sacrifice from the other,

depending on what the experiment requires. For example, a smaller I.D. is more efficient

for separation, but it has lower sampling capacity, resulting in signal saturation. Decreasing

the thickness of the packed column also decreases film thickness, improving resolution and

signal-to-noise ratio (S/N ratio); the disadvantage is that it also decreases sampling capacity.
49

3.6.3 Principles of GC as a Refinery Gas Analyzer (RGA)

The GC system for quantifying O2, CO2, CO, CH4, C2H4, etc. is a specific gas

chromatograph configured with three parallel channels which can analyze different species

in several columns simultaneously; it is used mainly in refinery gas analysis. After the

sampling injection, analytes are divided into three channels, one FID and two TCDs. Light

hydrocarbons from C1 to C6 are detected on the FID channel while permanent gases are

determined by a TCD. The second TCD is specifically designed for H2 detection, with only

N2 as the reference gas. In the valve system, two ten-port valves are used for looping prior

to sampling, back-flushing heavier components in lines [97]. Three six-port valves are

assigned different tasks before injection to each detector. PCM is the pressure control

module, which provides two channels of gas control. EPC is used to control the pressure

at the inlet section. The runtime event for all valves is described in Table 3.4.

V5

V4
V1

V3
V2

Figure 3.12: Plumbing diagram in GC-RGA. V1 and V5 are ten-port valves with

backflush function. V2, V3 are six-port valves with functions of column isolation to TCD,
50

column sequence reversal with backflush of the pre-column. V4 is the second six-port

valve with a loop for sampling upstream between V1 and V5 [97].

3.6.3.1 Analysis Sequence (Permanent and Light Gases Channel)

In diagram 3.12, a sampling gas mixture is imported from PCM B Channel one and

two, flowing through valve one, column one, column two, valve two, column three and,

eventually introduced to TCD. A simplified graph is shown in Fig. 3.4. The heaviest

species are trapped in column one and finally flushed away by the backflush function. O2

and H2 are light enough to be separate from the rest. The remaining light gases are trapped,

either in 4Ft Unibeads 1S or Molecular Sieve 5A, depending on their physical properties.

After approximately three minutes, valve one switches off and C3+ compounds are

backflushed away from column one, as shown in Fig. 3.5. In the meantime, valve two (also

called the isolation valve), is switched on to introduce ethane and carbon dioxide directly

to TCD. The purpose of isolation is to protect the Molecular Sieve column from carbon

dioxide, since it would be trapped and deactivate the column. Carbon monoxide and

methane are left in Molecular Sieve 5A. Finally, approximately five minutes after, valve

two is switched off so that carbon monoxide and methane elute from the molecular sieve

column to the TCD.

Valve 1 Valve 2 TCD

C2 + CO2, C2H6 CO, CH4, N2 O2, H2

2Ft Unibeads 1S 4Ft Unibeads 1S Mol Sieve 5A (Col. 3)


(Col. 1) (Col. 2)

Figure 3.13: Flowsheet of light hydrocarbon gases in valve one and two prior to

backflush process.
51

CO2
C2 CO, CH4 C2H6
+

2Ft Unibeads 1S 4Ft Unibeads 1S Mol Sieve 5A (Col. 3) TCD


(Col. 1) (Col. 2)

Figure 3.14: Flowsheet of valve two as isolation valve is switched on, valve one is off

with backflushing.

3.6.3.2 Analysis Sequence (Hydrocarbon Channel)

The hydrocarbon mixture separation process is simpler than the process using light

and permanent gases, As valve four switches on, samples in the loop from the ten-port

valve are introduced to DB-1 and HP-Al/S columns (Fig. 3.15). After a time period, C5

elutes to the HP-Al/S column where valve three is located. The purpose of valve three is to

make C6+ exit early as one peak on the chromatograph so that backflush and reverse flow

of DB-1 are implemented (Fig. 3.16).

FID

C1-C5 C6 +

25 m HP-Al/S 2m DB-1
(Col. 7) (Col. 6)

Figure 3.15 Flowsheet of hydrocarbons in valves three and four, before backflush
52

FID

C1-C5 C6 +

25 m HP-Al/S 2m DB-1
(Col. 7) (Col. 6)

Figure 3.16: Flowsheet of backflush and reverse flow as valve three is switched on.

Table 3.4: Runtime event for all valves in RGA.

Time Set point (min) Position On/Off


0.1 Valve 1 On
0.1 Valve 5 On
0.1 Valve 4 On
0.5 Valve 3 On
0.5 Valve 5 Off
1.0 Valve 4 Off
3.0 Valve 1 Off
4.20 Valve 2 On
5.0 Valve 2 Off
5.9 Valve 3 Off

3.6.4 Principles of GC/MS/FID/TCD

The other system used in this study was GC/MS/FID/TCD, for measuring levulinic

acid, EL, ethyl vinyl ketone and other oxygenated species. With this device, sampling in

lines were separate, with splitters dividing into three flows to MS, FID, and TCD from the

mainstream.

3.6.4.1 Mass Spectrometry Detector

The Mass Spectrometry Detector (MSD), shown in Fig. 3.17, coupled with a GC

system for species identification, is the most widely used detector. After separation from
53

the GC columns--and depending on their polarities and molecular weights--mixtures are

delivered into MSD and ionized, first with either electrical ionization (EI) or chemical

ionization (CI) under ultra-low pressure (1.33x10-5 Pa). However, EI is considered to be a

hard ionization method that forms fragments, and the signal on the mass spectrometer of

the parent compound is relatively small. All ions are introduced into the quadrupole ion-

trap mass analyzer for further separation. Eventually, species are magnified with a

multiplier and several mass-to-charge (m/z) ratios are presented as peaks on a spectrum.

With an already existing NIST library, each spectrum from GC/MS analysis is compared

with the database to determine the exact structure, name and molecular weight. Figure 3.18

shows the EL mass spectrum from the GC/MS system; smaller numbers are fragments

occurring after the ionization process.

Focal lenses

MSD
-
e
-
e Signal
-
e

GC column Ion trap analyzer

Electrical ionization

Figure 3.17: Simplified scheme of Mass Spectrometer Detector with electrical ionization

method as the ion source.


54

Figure 3.18: Mass spectrum of ethyl levulinate from NIST library.

3.6.4.2 Flame Ionization Detector and Effective Carbon Number

The flame ionization detector is designed mainly for organic compound detection.

With organic compound combustion in hydrogen flame, ions are generated and flame

intensity is proportional to the concentration of organic species. Figure 3.19 illustrates a

sample flow representing organic compounds passing through a hydrogen diffusion flame,

diluted in the air in which the unknown species are combusted, ionized and eventually

collected by an electrode [98]. The ion formation in the flames derives mostly from a

chemi-ionization of CHO that produces an electron, described below:

CH + O → CHO+ + e-

Based on this theory, different chemicals with a specific number of carbons and hydrogens

have different electrons that are generated after ionization; this unique response can be

interpreted as the effective carbon number (ECN). This method can be applied when

standard internal and external quantification methods are unavailable. However, a

heteroatom (neither H nor C), or the carbon nearby in an organic compound, often offers a

less than normal contribution to the response. For example, the carbon in a carbonyl group
55

makes zero contribution to the ECN and an oxygen atom in an ether has a negative

contribution. A reference method, clarifying the contributions of functional groups, is listed

in Table 3.5.

Charge collector

-
+

Hydrogen
Air
Sample flow

Figure 3.19: A basic design of flame ionization detector [98].

Table 3.5: Contributions to the effective carbon number [99].

Atom Functional group ECN contribution

C Aliphatic 1

C Aromatic 1

C Olefinic 0.95

C Carbonyl 0

C Carboxyl 0

O Primary alcohol -0.5

O Ether -1.0
56

3.6.4.3 Thermal Conductivity Detector

The thermal conductivity values of non-organic species’ detection, such as O2, N2,

CO, etc. are all different, so this material property can be implemented to construct a

detector for analyzing gas mixtures. When a species is introduced to a regime with

temperature gradients, conduction of heat takes place from the higher side to the lower

temperature points. In Eq. 3-7, with a known heat flow 𝑄, surface area 𝐴, temperature

points (𝑇1 , 𝑇2 ) and the distance 𝑥, thermal conductivity 𝜆 can be calculated and identified.

𝐴(𝑇1 −𝑇2 )𝜆
𝑄= (Eq. 3-7)
𝑥

In the design of TCD, several coiled filaments are mounted in the chamber to gain

maximum resistance. In the design of the Wheatstone Bridge, for example [96], the column

gas flow was measured, transformed to an electrical signal and compared with the reference

flow. The only disadvantage of TCD is that it is less sensitive and quantitative than FID

for good resolution of organic species analysis; for this reason alone N2, O2, CO, and CO2

were analyzed by TCD in this study.


57

Figure 3.20: Bridge circuit used in two-cell detector. R3 is used for sample flow detection

from column. R4 measures a reference flow from nitrogen or helium [96].


58

Chapter 4 JSR VALIDATION EXPERIMENT

Because the system was not validated in the CCRC, before the oxidation study of

ethyl levulinate (EL) began, the jet-stirred reactor model was tested.

Fuels with relatively well-developed mechanisms such as n-heptane, and iso-octane

(2,2,4-trimethylpentane) have been studied by Herbinet (2012 [50]), Zhang (2016 [100]),

Tsang in (2015 [101]) and Atef (2017 [102]), respectively; and they can be used to provide

validation of the setup. All experimental products were analyzed either through online

sampling in the gas chromatography-refinery gas analyzer (GC-RGA), or via syringe

injection to GC/MS/FID/TCD.

Three sets of data were obtained for the oxidation of n-heptane. For cyclohexene,

pyrolysis was studied because of its simple unimolecular decomposition reaction pathway,

forming ethylene and 1.3-butadiene. All fuels were tested under two conditions with the

sampling probe or the sampling probe with a pumping system. Results indicated that the

probe that included a pumping system best fitted previous data and simulation results. The

conclusion was that the residence time of the end-products in the sampling probe was the

main cause of the data discrepancies.

4.1 N-heptane Oxidation

N-heptane was selected because it is one of the primary reference fuels (PRF) for

determining research and motor octane number (RON, MON), and because it is a typical

diesel fuel. It has been thoroughly tested under a wide range of conditions in several

reactors and devices for proving low-temperature chemistry (in Supplemental Material),

which can be detected through temperature profiles by jet-stirred reactors [48, 50, 100,

103-105] or via ignition delay times with shock tubes apparatus [106-110] . The detailed
59

mechanism was developed and compiled by Curran in 1998 [111] and Zhang [112]. In this

study, n-heptane was tested using experimental parameters similar to Herbinet in Nancy

[50]. All parameters of the JSR in the literature are summarized in Table 4.1 and compared

with the facility at KAUST. Only one equivalence ratio was chosen (φ = 1) and repeated

under this condition.

Table 4.1: JSR experimental parameter comparison between Nancy and KAUST

JSR in Nancy JSR in KAUST


Pressure (Pa) 1.074 × 105 1.013 × 105
Residence time (s) 2 2
Equivalence ratio φ 0.5, 1.0, 2.0, 4.0 1.0
Temperature range (K) 500-1000 480-1100
Fuel concentration (%) 0.5 0.5
Reactor volume (cm3) 95 76
Number of preheating zones 2 1
Carrier gas He N2
Sampling method online and offline online and syringe injection

Nancy Apparatus

In the setup at Nancy, oxidation was under constant pressure, 1.074 × 105 Pa,

residence time was two seconds, temperature ranged from 500 to 1100K and equivalent

ratios of φ = 0.5, 1.0, 2.0 and 4.0; fuel was diluted in helium with 0.5% fuel concentration.

In the experimental apparatus at Nancy, two pre-heating systems were installed between a

vaporizer and a JSR: one system for preventing fuel condensation and the other for the pre-

heating zone, which is same as the set point of the furnace temperature. The injection

microprobe, upstream, and sampling probe, downstream, were installed on either side of

the JSR.
60

KAUST Apparatus

To repeat the experiment, several configurations were tested in order to acquire the

same results as the experiment in Nancy [113].

1. Without the sampling probe: experiments were conducted without a sampling

probe to test the significance of the probe. The injection probe was designed

with I.D. 1.6 mm, O.D. 2 mm, and a very small orifice with I.D. 0.7 mm made

of quartz. A glass finger, inserted with a thermocouple, was placed downstream

of the tube and connected to the JSR to measure temperature (Fig. 4.1). Syringe

injection was implemented to introduce the sample to the GC/MS/FID/TCD.

2. With sampling probe: a quartz sampling tube was installed next to the glass

finger (Fig. 5.1 (b)). The main reason for incorporating a sampling probe was

to reduce the residence time of end-products in the furnace, which has a similar

configuration as the injection microprobe. The volume flow rate at the outlet of

the sampling probe was measured at around 10 sccm (standard cubic

centimeters per minute). The residence time in the sampling probe was reduced

to 0.024 s, 1.2% of the residence time in the JSR.

3. With sampling probe and pumping system: a vacuum suction pump was

additionally attached at the exhaust pipeline of the GC system. The volume flow

rate of the sampling probe was improved to around 60 sccm, which exhibits

0.2% of the residence time in the JSR.


61

Pumping
device

Figure 4.1: Fuel injection probe I.D. 1.6 mm with 0.7 mm exit. From top to bottom: without

sampling probe; with sampling probe; with sampling probe and pumping device.

4.1.1 Oxidation Experiment of n-heptane

The results with/without a sampling probe are presented in Fig. 4.2, with the

experimental and simulation results from Zhang in 2016 [100]. The simulation was
62

performed using the perfect-stirred reactor module and transient solver program, with an

end-time of 20 s, using Chemkin-Pro software [114].

In examining the fuel consumption profile, note that at the beginning good

consistency is shown with experimental data from the literature [100], though all results

deviated from the modeling prediction. Alternative pathway reactions during low

temperature were well observed in both tests. With the installation of the sampling probe,

the temperature region (650 to 800 K), fits well with experiments from Zhang [100].

However, the results from the experiment without the sampling-probe dropped to only 0.2

to 0.3% of fuel remaining after heating to 700 K or higher. To explain this phenomenon:

When fuel and O2 exit from the reactor to the downstream port without the sampling probe,

the volume of the tube is 323 times larger than the volume of the sampling probe and

exhibits 3.37 s residence time after leaving the reaction zone. Compared with the sampling

probe, this device can be taken as a plugged flow reactor with a higher residence time.

There is an additional reactive temperature zone in which reactions continue to occur,

resulting in 40% increased fuel consumption. At this point, experiments without the probe

ceased, owing to poor temperature profiles, which began at 700 K.

Next, under a high-temperature regime (900 to 1000 K), the results from this

apparatus (with the sampling probe and without a pumping device) showed higher fuel

consumption and more O2 remaining than the results in the literature. It can be assumed

that residence time in the sampling probe (1.8 s) was still too long, causing some fuel to

undergo pyrolysis. An insufficient premixing zone was also suggested as the cause of

excessive fuel consumption. As a result, another set of experimental data was collected,
63

combining fuel with O2 and N2 prior to the premixing zone; the results were identical to

the previous apparatus (separating fuel and O2 at the entrance of microprobe).

Finally, a pumping device was introduced to achieve lower residence times in the

sampling probe. The volume flow rate was increased to 60 sccm, six times higher than

previously. Under this circumstance, residence time in the sampling probe was reduced to

15% of the residence time of the mixture in the JSR.

Updated data on n-heptane, O2, CO, and C2H4 appears in Fig. 4.3 and is compared

with the literature [100] and the data on the probe without the pumping device. Even though

both species were under-predicted by the mechanism from Zhang [100], results from this

study have good trajectory, following the published results. More O2 was consumed than

in the previous test at high temperature (> 900 K) and n-heptane concentration reached

0.5% between 700 and 750 K. In conclusion, decreasing residence time in the sampling

probe has a significant impact on n-heptane and O2 consumption profiles versus different

temperatures.
64

0.060
0.005
0.055

0.004 0.050

0.045

Mole fraction
Mole fraction

0.003
0.040

0.002 0.035

Zhang_2016 0.030 Zhang_2016


0.001
without probe without probe
with probe 0.025 with probe
Zhang_2016
0.000 Zhang_2016
0.020
500 600 700 800 900 1000 500 600 700 800 900 1000
Temperature (K) Temperature (K)

0.005 0.020
Zhang_2016 Zhang_2016
0.018
without probe without probe
0.004 with probe 0.016 with probe
Zhang_2016 Zhang_2016
0.014
Mole fraction

Mole fraction
0.003 0.012

0.010
0.002
0.008

0.006
0.001
0.004

0.002
0.000
0.000
500 600 700 800 900 1000 500 600 700 800 900 1000
Temperature (K) Temperature (K)

Figure 4.2: Experimental results for n-heptane oxidation with 0.5% fuel concentration, 1

atm, 2 s residence time and equivalence ratio 1. Figure 4.1-a without sampling probe, 4.1-

b with sampling probe, respectively. Configurations for blue and red dots in portrait

position. From left to right, from top to bottom: n-C7H16, O2, CO2, and CO.
65

0.060
0.005
0.055

0.004 0.050

0.045

Mole fraction
Mole fraction

0.003
0.040

0.002 0.035

0.030 Zhang_2016
0.001 Zhang_2016
with probe with probe
with probe & pumping 0.025 with probe & pumping
Zhang_2016 Zhang_2016
0.000 0.020
500 600 700 800 900 1000 500 600 700 800 900 1000
Temperature (K) Temperature (K)
0.020 0.006
Zhang_2016 Zhang_2016
0.018
with probe with probe
0.005
0.016 with probe & pumping with probe & pumping
Zhang_2016 Zhang_2016
0.014 0.004
Mole fraction

Mole fraction
0.012

0.010 0.003

0.008
0.002
0.006

0.004 0.001
0.002
0.000
0.000
500 600 700 800 900 1000 500 600 700 800 900 1000
Temperature (K) Temperature (K)

Figure 4.3: Comparison of results of n-heptane, O2, CO, and C2H4 (top to bottom, left to

right) with different experimental configurations.

4.2 Cyclohexene Pyrolysis

Cyclohexene pyrolysis has a very simple unimolecular decomposition pathway, the

Retro-Diel-Alder reaction, producing only ethylene and 1,3-butadiene. [101]:

C6H10 → C2H4 + C4H6

Its rate coefficient, adopted in this study, was obtained from the low-pressure flow

experiment: k(T) = 1015.15exp(-33,233/T) [115]. This basic decomposition was observed

in the JSR experiment in this study in order to check the influence of the pumping system.

In Fig. 4.4, it is apparent that cyclohexene was consumed faster at high-temperatures (>
66

850 K) and products were formed earlier than predicted. Since no premixing problem

existed in the pyrolysis study, the only possibility for improvement was to install the

pumping system. As a result, another set of experiments were conducted; the results are

described in Fig. 4.5. Eventually, species’ profiles matched the simulation results well. The

necessity to avoid further reactions in the sampling tube was apparent.

0.0014
C2H4 with probe
0.0010
0.0012 C4H6 with probe

0.0008 0.0010
Mole Fraction

Mole fraction
0.0008
0.0006
0.0006

0.0004 0.0004

0.0002
0.0002
0.0000
0.0000 C6H10 with probe
-0.0002
800 900 1000 800 900 1000
Temperautre (K) Temperature (K)

Figure 4.4: Experimental results of cyclohexene pyrolysis in JSR. Only ethylene and 1,3-

butadiene quantified under 1 atm, 0.1% fuel and 1 s residence time from 750 to 1025 K.

0.0014
C2H4 with probe & pumping
0.0010
0.0012 C4H6 with probe & pumping

0.0008 0.0010
Mole Fraction

Mole fraction

0.0008
0.0006
0.0006

0.0004 0.0004

0.0002
0.0002
0.0000
0.0000 C6H10 with probe & pumping
-0.0002
800 900 1000 800 900 1000
Temperautre (K) Temperature (K)

Figure 4.5: Experimental results of cyclohexene pyrolysis with sampling probe and

pumping device configuration with identical conditions and modeling as Fig. 4.4.
67

4.3 Iso-octane Oxidation

The iso-octane oxidation mechanism was studied in the JSR in several works of

literature [48, 116, 117]. In Dagaut’s article [48], iso-octane was first full-tested in a JSR

under 10 atm, mean residence time 1 s, 0.1% fuel with equivalent ratios φ = 0.3, 0.5, 1.0

1.5. In this experiment, iso-octane was oxidized at temperatures from 600 to 1000 K, 1 atm,

φ = 1 and 0.5% fuel with 2 s mean residence time. The main hydrocarbon species detected

in GC-RGA FID were CH4, C2H4, C3H6, i-C8H18 and O2, CO in TCD. The kinetic and

thermodynamic files for simulation were provided by Atef et al. [102].

Simulation and experimental results are compared in Fig. 4.6. The fuel

consumption profile was acquired from 600 to 1000 K to examine negative temperature

coefficient (NTC) behavior (600 to 700 K). The set of experimental results without the

pumping system did not match the model prediction, especially under high temperature

conditions. O2 was over-consumed while CO was over-produced.

However, the experimental results with the pumping system agreed well with

simulation results, improving the species concentration profiles; but the pumping system

did not greatly improve the effect on the propylene profile and no difference was detected

in the ethylene production profile under the intermediate temperature range (800 to 900 K).
68

with probe 0.0040 with probe


0.020 with probe & pumping
with probe & pumping
0.0035 CH4
iso-octane
CO C3H6
0.015 0.0030

Mole Fraction
Mole fraction
0.0025

0.010 0.0020

0.0015
0.005
0.0010

0.0005
0.000
0.0000

600 700 800 900 1000 600 700 800 900 1000
Temperature (K) Temperature (K)

0.070
with probe
0.0015
0.065 with probe & pumping
C 2 H4
0.060
Mole Fraction

Mole Fraction
0.0010
0.055

0.050 0.0005

0.045
with probe
with probe & pumping 0.0000
0.040
O2
600 700 800 900 1000 600 700 800 900 1000
Temperature (K) Temperature (K)

Figure 4.6: JSR species concentration profiles (points) and simulations, using updated

mechanism from Atef et al. [102] for 0.5% fuel and 2 s residence time under atmospheric

pressure.
69

Chapter 5 ETHYL LEVULINATE MODELING AND EXPERIMENT

The primary goal of this study is to optimize the kinetic model of ethyl levulinate

using oxidation studies data obtained from the validated JSR setup. The mechanism was

developed by Dr. Stephen Dooley and the thermochemistry calculation was performed by

Dr. Manik Kumer Ghosh. In addition to utilizing GC-RGA for sample detection and

quantification, some experiments were conducted via syringe injection to GC, coupled with

a mass spectrometry detector (MSD), a flame ionization detector (FID) and a thermal

conductivity detector (TCD) under several equivalence ratios from 0.5 to 2, N2 as carrier

gas, 0.5% fuel at 1.013 × 105 Pa from 600 to 1000 K.

5.1 Ethyl levulinate Oxidation Mechanism

Based on the fuel structure, EL contains both a ketone [118, 119] and an ester

functional group [69, 74] resulting in markedly different results from conventional

paraffinic hydrocarbons and other oxygenates. The purpose of developing the EL kinetic

model is to aid understanding the use of this potential biofuel. The EL kinetic mechanism

is built on the top of a gasoline surrogate mechanism [120], composed of C0-C4; primary

reference fuels: n-heptane and iso-octane; 1-hexene as olefin and toluene as an aromatic

compound. The chemistry of several species such as HCOOC2H5, C2H3COOH, and

C2H3COCH3 were also included in the model.

5.1.1 Kinetic Model

The elementary reaction pathways, their rate constant expressions and the

thermochemical properties of the radicals and molecules pertinent to EL combustion, have

been prescribed, considering the best knowledge available for each discrete reaction center,

as studied in other simpler systems. In general, for the reaction centers comprising the
70

methyl ketone functionality, reaction parameters have been assigned through analogy to

those of methyl ethyl ketone. Reaction centers comprising the ethyl ester functionality,

analogies to those of the well-studied ethyl propanoate (EP) have been made, principally

considering the modelling works of Serinyel et al. [118] and Metcalfe and co-workers [27],

respectively. Bond dissociation energy (BDE) of EP is one of the most critical factors for

kinetic and thermodynamic parameters calculation. Depicted in Fig. 5.1 (a), the carbon-

hydrogen BDE is generally weaker at the α-carbon position in the (C*O)-alkyl moiety of

the molecule. Moreover, as C-H bond breaks, the whole molecule is stabilized by the

carbonyl group resonance, resulting in a decreased reaction rate with other radical species,

delineated in Fig. 5.1 (b).

97.3
85.8
101.1 100.2 89.4
91.8 86.6 102.1
α
94.2

Figure 5.1: (a) Bond dissociation energy of ethyl propanoate (EP) (kJ/mol). Black

numbers are C-C or C-O BDE, red numbers are C-H BDE [29]. α-carbon is designated to

the carbon next to carbonyl group (C(=O)). (b) Radical stabilization by the carbonyl

group.

As a radical site at α-carbon forms, the radical is stabilized due to the carbonyl

group resonance, slowing down the relative reaction rates.


71

The appropriateness of the analogies imposed above is supported by a similarity in

the thermochemical environment local to each reaction center, as depicted in Fig. 5.2. The

contributions of the ketone group are adopted from those reported by Hudzik et al. [121]

and the ethyl ester contributions are consistent with the finding of El-Nahas et al. [122] for

the case of EP. These calculations result in a set of bond dissociation energies for EL, as

shown in Fig. 5.2.

(a)
93.6 102.8
82.2 95.2 88.6
84.0 79.9 99.0 90.8
97.0 90.9 97.5

(b)
97.3
101.1 100.2 89.4

85.8 91.8 86.6 102.1


94.2

(c)

101.5
95.6
90.8

Figure 5.2: Chemical structure and bond dissociation energy (BDE) of (a) ethyl levulinate

calculated by Dooley et al. [123], (b) ethyl propanoate (EP) [122] and (c) methyl ethyl

ketone (MEK) [121]. Red numbers represent C-H BDE, black numbers are C-C or C-O

BDE. (Unit: kcal/mol)

To build a complete kinetic mechanism for EL, Dooley suggested that several sub-

mechanisms should be served and assembled. The basic chemistry is given by a C0 to C4

mechanism. Moreover, from C5 to C8, hydrocarbon chemistry is provided from the

primary reference fuel (linear and branched saturated hydrocarbons), olefins (1-hexene)
72

and aromatics (toluene) from surrogate gasoline mechanism [120]. Furthermore EL and

intermediate reactions are included, adopted from various sub-models. Temperature

dependent reaction rates are expressed in the model using the modified Arrhenius equation:

k(T) = A.Tn.exp(–Ea/RT), where A is the pre-exponential collision frequency factor, T is the

temperature, n is the fitting parameter, Ea (cal mol-1) is the activation energy, and R

represents the universal gas constant. In most temperature ranges encountered in

combustion, temperature dependence can adequately be integrated into A factor (with n =

0). However, additional variation is considered for some over-range reactions in which an

appropriate value of n accurately fits the available experimental data [124]. Reaction

pathways have been prescribed, generally according to the protocols of Curran et al. [125].

The high-temperature reaction classes considered in the development of this model are

listed below.

1. EL unimolecular decomposition. Several decomposition pathways should be

involved to produce smaller radicals. In the mechanism, radical decomposition rate

constants are estimated in the reverse exothermic direction (P+Q => R), allowing

the forward rate constant (R=> P+Q) to be calculated from the equilibrium constant

in the thermodynamic file, and most reverse rate coefficients are assigned to 3×1013

(cm3 mol-1 s-1).

Ethers, esters, and alcohols preferably undergo molecular elimination reactions

[126-128] during oxidation or pyrolysis. Metcalfe et al. [27] presented the high-

pressure limit rate for the EP molecular elimination, mentioned in Chapter 2.3.2:

EP  C2H5COOH + C2H4 (R1)


73

with a rate expressions k(T) = 1.6×1013exp(-50,000/RT) s-1. Westbrook et al. [128]

found similar rate expressions with only A factor differences of 2.0×1013 and

1.0×1013 for ethyl acetate and ethyl formate, respectively. In EL, the ethyl ester is

also concluded to undergo elimination, forming levulinic acid (LA) and ethylene

(R2), as the dominant reaction. In this study, the EP rate coefficient was chosen for

simulation in the kinetic mechanism.

2. Hydrogen abstractions from fuel. This class includes all rate coefficients for

abstraction of H atoms from the fuel by small radicals such as H, OH, CH3, CH3O,

CH3O2, HO2, O and O2. Dooley adopted most of the rate coefficients in this class

from EP and MEK, with minor modification to the A factor or activation energy Ea.

3. Fuel radical decompositions. After H abstraction, five fuel radicals proceed, by

further decomposition, to produce smaller species, usually olefins and other

radicals at high temperatures. Along with the translation of this model, most rate

coefficients were developed using both the EP and MEK models. (Note that C-O

beta-scission reactions are also taken into account.)

4. EL radical isomerizations. All rate coefficients (relative to isomerization of the

five fuel radicals) are again assigned as identical, as 1×1013. Actual rates are

determined by thermochemistry of fuel radicals.

5. Ethyl formate formation and consumption. By breaking the C-C bond between

the ketone and carbonyl group, ethyl formate (C2H4OC(O)H) becomes a potential

intermediate and should be considered; its sub-mechanism was developed from

Westbrook et al.[128].
74

6. Levulinic acid (LA) relative reactions. As one of the major products from EL

decomposition reactions, some LA relative reactions, such as hydrogen abstraction

reactions, are taken from methyl cyclohexane sub-mechanism [129] with minor

adjustments. Beta-scission reactions are using analogy to either EP or MEK sub-

mechanism.

7. Ethyl propanoate and propenoate submodel. As EL decomposes to smaller

radicals, or stable species, some structures are similar to that of EP, or unsaturated

compounds like ethyl propenoate. The decomposition reactions relating to these

specific compounds are taken from analogous reactions in the EP sub-model [27].

8. C2H3COCH3 consumption. Since methyl vinyl ketone has been predicted and

verified as one of the most abundant products from EL oxidation, its further

decomposition reactions (including hydrogen abstractions by small radicals or

unimolecular decomposition), are also considered. Rates are taken from the three-

pentanone oxidation model [130].

5.1.2 Ethyl Levulinate Thermochemistry

The thermodynamic parameters were taken mainly from the surrogate fuel [120].

Regarding the thermochemistry of ethyl levulinate (EL, CH3C(O)CH2CH2C(O)OCH2CH3),

levulinic acid (LA, CH3C(O)CH2CH2C(O)OH) and five primary fuel radicals, Dr. Manik

Kumer Ghosh performed the calculations for EL, LA and other relative radicals in the

GAUSSIAN 09 package [131]. The geometries were optimized with B3LYP/6-31G(d,p)

theory. The thermochemistry of all species produced by the homolytic cleavage of EL and

their consumption products, not described in the EL sub-model, were estimated by group

additivity, employing THERM software [132]. This method has been used for over 20
75

years; it is much faster than the higher-level ab initio method and it provides a good

estimation.

Overall simulating files were accomplished, including the kinetic mechanism and

thermochemistry; modeling computations in this study were carried out using the perfect-

stirred reactor code in CHEMKIN PRO [114].

5.2 Ethyl Levulinate Oxidation Experiment

Before systematically quantifying the EL oxidation experiment at different

equivalence ratios, several intermediates were detected using GC/MS/FID/TCD; the

spectrum appears in Fig 3.18. Some unknown species were identified by MS detector,

showing over 85% mass spectra similarity with the NIST library: ethanol C2H5OH, methyl

vinyl ketone C4H6O, ethyl acrylate C5H8O2, angelica lactone C5H6O2, ethyl levulinate

C7H12O3 and levulinic acid C5H8O3. Thion [44] performed methyl levulinate reaction

pathway theoretical calculations and predicted the production of methyl vinyl ketone and

methyl acrylate. Interestingly, methyl vinyl ketone and ethyl acrylate were also formed

from the decomposition of EL. These two compounds, and levulinic acid, were the highest

peaks recorded by GC/MS analysis after EL oxidation (Fig. 5.3), excepting ethyl levulinate

(peak F) and nitrogen (peak A).


76

GC-MS @ 850 K
4 F

3
Int. (x107)
2 G
C

1
A
B D
E
0

6 8 10 12 14 16 18 20 22 24
Time (min)

Figure 5.3: GC-MS spectrum of EL oxidation under 850K in JSR unknown species

identified with MS detector. From A to G: A: nitrogen; B: ethanol; C: methyl vinyl

ketone; D: ethyl acrylate, E: angelica lactone, F: ethyl levulinate and G: levulinic acid.

The JSR was next used to validate the mechanism under 1 bar, 600 to 1000 K, 0.5%

fuel, 2 s residence time and equivalence ratios from 0.5 to 2. The boiling point of EL is

around 206 oC and the vaporizer was heated to 240 oC. Both upstream and downstream

transport lines were heated to 230 oC. The experimental results of the EL profile were

measured with GC/MS/FID/TCD with a sampling probe and syringe injection method (Fig.

5.4); it has not as yet been detected by GC-RGA, so online analysis is not available. The

remainder of the species: ethylene C2H4, methane CH4, hydrogen H2, oxygen O2, carbon

monoxide CO and carbon dioxide CO2 were all analyzed and quantified by online analysis.

The pumping system has not yet been implemented with this fuel.
77

EL exhibits a similar consumption profile to ethyl propanoate as the concentration

drops significantly at 775 K and is fully consumed after 900 K; but the prediction of

ethylene is underestimated and more ethylene was detected from 700 to 850 K. Hydrogen,

methane and carbon monoxide consumption profiles are acceptable for φ = 0.5 and 1.

However, oxygen is always under-predicted and the estimation is lower with increasing

equivalence ratios. For CO and CO2, experimental signals are nearly twice as great after

900 K, when φ is equal to 2. Similar tendencies was also observed in iso-octane

experiments, as underestimation of CO, CO2 occurred when the experiment was conducted

without a pumping system. In summary, some rate coefficients in the original kinetic

mechanism should be altered, especially those relating to ethylene production.


78

0.006
EL
0.08
C 2H 4
0.005
CH4
H2 0.06
0.004

Mole fraction
Mole fraction
O2
0.003
0.04 CO2
0.002 CO

0.02
0.001

0.000 0.00

600 700 800 900 1000 600 700 800 900 1000
Temperature (K) Temperature (K)

0.006
EL
C2H4 0.04
0.005
CH4
H2 0.03
0.004

Mole fraction
Mole fraction

O2
0.003 CO2
0.02
CO
0.002
0.01
0.001

0.000 0.00

600 700 800 900 1000 600 700 800 900 1000
Temperature (K) Temperature (K)

0.005 0.020

0.004
0.015
EL
Mole fraction

Mole fraction

0.003 C 2H 4 O2
CH4 0.010 CO2
0.002 H2 CO

0.005
0.001

0.000 0.000

600 700 800 900 1000 600 700 800 900 1000
Temperature (K) Temperature (K)

Figure 5.4: Temperature profiles of ethyl levulinate (EL) oxidation, with 0.5% fuel, 1 bar,

2 s residence time and equivalence ratio from top to bottom: φ= 0.5, 1, 2.


79

5.3 Kinetic Model Adjustment

Comparing the model validation against experimental results showed some

disagreements, therefore further adjustments and improvements to the chemical kinetic

model must be made, especially in fuel unimolecular decomposition, H-abstraction and

fuel radical derived carbon-carbon (C-C) and carbon-oxygen (C-O) scissions reactions.

The kinetic mechanism was updated by applying new rate coefficients from another

member of the levulinic family, methyl levulinate (ML, CH3C(*O)CH2CH2C(O*)OCH3).

Thion et al. [44] calculated the theoretical enthalpy of formation of different species

and radicals derived from ML. Thereafter, rate coefficient computation was performed for

two classes applied in this model: H atom-abstraction from the ML by OH and CH3 radicals

and fuel radical derived C-C and C-O scissions. In addition, rate coefficients for fuel

unimolecular decomposition were replaced with analogous rates from methyl butanoate

(MB) [28], ethyl propanoate (EP) [27] and methyl ethyl ketone (MEK) [133] mechanisms.

5.3.1 H-abstraction from The Fuel

In the fuel radicals shown in Fig. 5.5 (EL1J, EL3J, EL4J, EL6J, EL7J) only three

rate coefficients were adopted from ML calculation [44]. The rate coefficients are

described in Table 5.1, with their source. It was originally assumed that EL → EL1J has

the same rate coefficient as EL → EL6J and EL → EL3J has the same rate coefficient as

EL → EL4J. Both sites (C1 and C6) have similar C-H bond dissociation energy (BDE),

leading to the identical rate coefficient of EL1J and EL6J. However, C3 and C4 sites were

assigned the same rate coefficients, even though C4 has 2.7 kcal/mol higher BDE than C3.

It can also be observed in the EP oxidation model [27] that C2-H is 3.1 kcal/mol BDE

lower than CE-H, but is assigned the same rate coefficient. The BDE order (smallest to
80

largest) of EL is C3-H < C4-H < C1-H < C6-H < C7-H. To refine the model, EL + R ↔

EL1J* + RH rate coefficient was replaced with that of ML + R ↔ MLR 6 + RH, EL + R

↔ EL3J* + RH with ML + R ↔ MLR4 + RH and EL + R ↔ EL4J* + RH with ML + R

↔ MLR3 +RH. By plotting the rate coefficients against temperature (Fig. 5.7) the updated

rate coefficient order from highest to lowest is: EL3J* > EL6J > EL4J* > EL1J* > EL7J,

compared with the previous order: EL3J = EL4J > EL6J = EL1J > EL7J. By following the

updates it can be seen that EL6J forms more easily than EL4J and EL1J. However this

finding is not consistent with the BDE order, as noted above; therefore, the only change

made was to increase the rate coefficient of EL3J, to make the order fit with C-H BDE,

which gives: EL3J* > EL4J > EL1J = EL6J > EL7J. In conclusion, only EL + R ↔ EL3J*

+ RH, using ML calculations, was altered in the fuel H-abstraction reaction class in the

model.
81

EL EP
1

4 3 7 1 E
5
1 2 3
3 6 2 M

EL1J EL3J EL4J

EL6J EL7J

ML6R ML4R ML3R

Figure 5.5: Radicals or fuels nomenclature of ethyl levulinate, ethyl propanoate, and

methyl levulinate.

Table 5.1: Ethyl levulinate and ethyl propanoate mechanism rate coefficients for H

abstraction by OH radical; cm3 mol-1 s-1 cal-1 units.

Reaction A (cm3 mol-1 s-1) n Ea (cal mol-1) Comments

EL + OH → EL1J + H2O 2.34E+07 1.61 -35 same as EL6J


82

EL + OH → EL1J* + H2O 4.06E+03 2.84 65 analogy to ML6R [44]

EL + OH → EL3J + H2O 5.73E+10 0.51 63 same as EL4J

EL + OH → EL3J* + H2O 3.25E+03 2.89 -1,500 analogy to ML4R [44]

EL + OH → EL4J + H2O 5.73E+10 0.51 63 EP model [27]

EL + OH → EL4J* + H2O 2.87E+02 3.06 -2,320 analogy to ML3R [44]

EL + OH → EL6J + H2O 2.34E+07 1.61 -35 EP model [27]

EL + OH → EL7J + H2O 1.76E+09 0.97 1,586 EP model [27]

EP + OH → EP2J + H2O 1.15E+11 0.51 63 BDEC2-H = 94.2 [122]

EP + OH → EPEJ + H2O 1.15E+11 0.51 63 BDECE-H = 97.3 [122]

* parameters from methyl levulinate theoretical kinetic study.

1013
EL1J, EL6J
EL3J, EL4J
EL7J
EL1J*
EL3J*
k/cm3 mol-1 s-1

EL4J*

1012

EL+OH• = ELRJ+H2O

1011
0.8 1.0 1.2 1.4 1.6
1000 K/T

Figure 5.6: Rate coefficient comparison of H-abstraction from fuel. Solid line: original

rates adopted from other molecules; dashed line: rates from the methyl levulinate

theoretical kinetic study [44].


83

5.3.2 Fuel Radical Derived C-C and C-O Scissions

EL1J, EL3J, EL4J, EL6J and EL7J, are the fuel radicals which are the main

precursors of several intermediates, such as methyl vinyl ketone and ethyl acrylate, via C-

C or C-O beta-scission. The original rate coefficients are analogous to EP [27] and MEK

[133] models. To update the mechanism--particularly for EL1J, EL3J, EL4J radicals--their

C-C beta scission reactions were replaced with those calculated for ML, giving the detailed

rate coefficients delineated in Table 5.2. Comparing the rates of all decomposition

pathways shown in Fig. 5.7, coefficients were reduced in all cases. These rate coefficients

from ML calculations are all included in the new mechanism.

Table 5.2: Rate coefficients for fuel radical derived C-C and C-O scissions, only EL1J,

EL3J and EL4J were altered.

Reaction A (cm3 mol-1 s-1) n Ea (cal mol-1) Sources

EL1J → CH2CO + EP3J 1.00E+14 0.0 3.10E+04 Acetone

EL1J* → CH2CO +EP3J 3.50E+11 0.58 3.52E+04 Analogy to MLR6

EL3J → CH3 + COCHCH2CO2CH2CH3 1.00E+14 0.0 3.10E+04 Acetone

EL3J → C2H3COCH3 + EFF 8.83E+15 -0.67 3.56E+04 EP [27]

EL3J* → CH3 + COCHCH2CO2CH2CH3 8.15E+12 0.3 4.15E+04 Analogy to MLR4

EL3J* → C2H3COCH3 + EFF 4.29E+12 0.3 3.54E+04 Analogy to MLR4

EL4J → CH3CO + EP1D 1.00E+11 0.0 8.42E+03 3-pentanone

EL4J → CH3COCH2CHCO +C2H5O 6.82E+16 -1.34 2.69E+04 3-pentanone

EL4J* → CH3CO + EP1D 6.78E+12 0.22 2.38E+04 Analogy to MLR3

EL4J* → CH3COCH2CHCO +C2H5O 8.45E+11 0.51 4.89E+04 Analogy to MLR3


84

108

106

104

k/ s-1
102

100

10-2 EL1J C2-C3 scission


EL1J* C2-C3 scission
0.8 1.0 1.2 1.4 1.6
1000 K/T

1010

108

106

104
k/ s-1

102

100

10-2 EL3J C1-C2 scission


EL3J* C1-C2 scission
10-4 EL3J C4-C5 scission
EL3J* C4-C5 scission
10-6
0.8 1.0 1.2 1.4 1.6
1000 K/T

1010

108

106

104
k/ s-1

102

100

10-2 EL4J C2-O3 scission


EL4J* C2-C3 scission
10-4 EL4J C5-O3 scission
EL4J* C5-O3 scission
0.8 1.0 1.2 1.4 1.6
1000 K/T
85

Figure 5.7: Rate coefficient comparison of EL3J decomposing to smaller species via

beta-scission at different C-C bond positions.

5.3.3 Fuel Unimolecular Decomposition

Concerning fuel unimolecular decomposition reaction class (via C-C and C-O),

original rate coefficients were estimated for radical re-combination; subsequently

decomposition reactions can be determined from the equilibrium constant from the

thermochemistry. These rates can be improved further by comparison to three other

oxidation kinetic models: MEK [133] EP [27] and MB [28]. First, the C-C BDE order,

from lowest to highest, is C3-C4 < C2-C3 < C1-C2 < O3-C6 < C6-C7 < C4-C5 < C5-O3;

the updated model is principally constructed according to this order. For C1-C2 and C2-

C3 bond breaking, rate coefficients were adopted from the MEK model [133]. The C3-C4

bond breaking rate was taken from the methyl butanoate model, and O3-C6, C6-C7, C4-

C5, C5-O3 unimolecular decomposition rate coefficients were from the EP model [27].

These rate coefficients are compared and shown in portrait mode in Fig. 5.8. It is interesting

that the rate coefficient for C2-C3 is much higher, giving the order: C2-C3 >> C3-C4 >

C1-C2 > O3-C6 > C4-C5 > C6-C7 > C5-O3. Another unexpected finding is that the order

of C4-C5 and C6-C7 is contradictory to the C-C BDE order. This unusual occurrence can

also be observed in the EP oxidation model [27]; therefore these rate coefficients were not

adjusted. To solve the abnormal C2-C3 rate coefficient, they were given the same rate

coefficient as C1-C2 since their C-C BDE are similar (82.2 and 84.0 kcal/mol, respectively),

with the same treatment as the rate coefficients of EL1J and EL6J formation.

Table 5.3: Rate coefficients for EL unimolecular decomposition reactions.


86

Reaction A (cm3 mol-1 s-1) n Ea (cal mol-1) Bond position

EL → ELCOJ + CH3 4.04E+90 -21.9 1.22E+05 C1-C2, MEK [133]

EL → CH3CO + EP3J 1.78E+19 -0.73 8.19E+04 C2-C3, MEK [133]

EL → CH3COCH2 + CH2CO2CH2CH3 3.80E+16 0 8.17E+04 C3-C4, MB [28]

EL → CH2CH2COCH3 + C2H5OCO 2.63E+27 -3.23 9.47E+04 C4-C5, EP [27]

EL → CH3COCH2CH2COJ + C2H5O 1.65E+24 -2.04 1.00E+05 C5-O2, EP [27]

EL → CH3COCH2CH2CO2J + C2H5 5.73E+25 -2.76 9.21E+04 O2-C6, EP [27]

EL → CH3 + ELME2 3.39E+21 -1.58 9.21E+04 C6-C7, EP [27]

108
105
102
10-1
10-4
k/ s-1

10-7 C1-C2
C2-C3
10-10
C3-C4
10-13 C4-C5
10-16 C5-O3
O3-C6
10-19
C6-C7
10-22
0.8 1.0 1.2 1.4 1.6
1000 K/T

Figure 5.8: Rate coefficient comparison of EL unimolecular decomposition, including C-

C and C-O bond scission.


87

5.3.4 EL Concerted Elimination Reaction

Note that the rate coefficient of EL = levulinic acid + C2H4 was originally adopted

from ethyl propanoate [27]: k(T) = 1.6×1013exp(-50,000/RT). Another rate coefficient was

measured by AlAbbad [45] and calculated from a series of shock tube experiments.

Ethylene concentration was monitored using CO2 gas laser absorption technology. Under

1 bar condition, the rate coefficient is expressed as: k(T) = 3×1014exp(-54,574/RT). The

fitting was adjusted and used in the model. The rate coefficients from [27] and [45] are

compared in Fig. 5.9 in the range of 600 K to 1000 K. Two rate-coefficient expressions are

similar, with a cross point around 750 K. Even under lowest and highest temperatures, the

rate difference is no greater than an order of magnitude; yet this minor discrepancy may

have a significant effect on the prediction of ethylene production.


88

102

101

Rate Coefficient (s-1)


100

10-1

10-2

10-3

10-4

10-5
Metcalfe et al. (1.6x1013exp (50,000/RT))
10-6 AlAbbad et al. (10-41.79xT-8.20exp (33,777/T))
600 700 800 900 1000
Temperature (K)

Figure 5.9: Rate coefficient comparison in two different studies. Black line is EP

oxidation mechanism [27], used in this study. Red line is shock tube experimental results

[45]

5.4 Simulation Results from Updated Kinetic Mechanism

Performance of the original and updated versions are depicted in Fig. 5.10; the

disparity is in the ethylene production profile. The updated version shows that the ethylene

fits well using the updated rate coefficients, although the EL is consumed earlier. The fuel

consumption rate was enhanced and the fuel concentration was over-predicted at

stoichiometric and rich conditions. The prediction for other species was similar to the

original simulation results.

The reaction pathways of ethyl levulinate were performed and described in Fig.

5.11. The analysis is taken under stoichiometric conditions, atmospheric pressure, with

temperature at 800 and 1000 K. At 800 K the fuel is consumed by 20%, the temperature
89

chosen for analysis. EL is mostly consumed by a concerted elimination reaction and over

80% of the fuel decomposed to LA and ethylene. Approximately 10% of fuel is consumed

by H abstraction reactions, mostly by OH and H radicals attacking C3 and C4 sites; this is

reasonable as C3 and C4 sites have the weakest C-H BDE (Fig. 5.2). The subsequent

pathways of LA undergo H abstraction, followed by the production of CH3CO and

propenoic acid. However, EL radicals proceed via beta-scissions to smaller, stable and

radical species, such as methyl vinyl ketone and ethyl acrylate, observed in GC-MS in Fig.

5.3. At the higher temperature of 1000 K, fuel is almost completely consumed and

undergoes further H abstraction. With EL3J there is enough energy to overcome the

activation energy barrier of C-C bond breakage, reducing the importance of isomerization

from EL3J to EL4J. From the analysis it is evident that EL unimolecular decomposition

dominates the consumption of EL at all temperature ranges.

The sensitivity analysis coefficient of 0.5 % fuel under two temperature conditions

(800 and 1000 K, φ = 1) is shown in Fig. 5.12. The sensitivity coefficients for different

reactions represent the percentage change of the species concentrations, relative to the

baseline concentration, in terms of the dependence of the reaction parameters. The positive

coefficient denotes a rate coefficient that promotes fuel formation while the negative

coefficient indicates fuel consumption. As expected, fuel is very sensitive to the concerted

elimination reaction (EL = C5H8O3 + C2H4) under both temperatures, and causes 80% fuel

consumption at all temperature, as seen in the reaction pathway analysis. Also, chain

termination reactions (2CH3 = C2H6, CH3 + HO2 = CH4 + O2 and hydroperoxyl radical

formation) usually inhibit the consumption of EL, decreasing the overall reactivity;

however, chain branching reactions to produce a hydroxyl radical (H2O2 = 2OH and H +
90

O2 = O + OH) exhibit a positive effect on fuel consumption, promoting overall reactivity

under both 800 and 1000 K, respectively. At 800 K the occurrence of H-abstraction

reactions of EL, such as EL + HO2 and EL + CH3O2, forming fuel radicals EL3J, EL4J,

predominate fuel consumption as well. When the temperature rises to 1000 K, the rate

coefficients of the basic chemistry of small radicals become more sensitive to fuel

consumption. (Note that the importance of ethylene chemistry: C2H4 + OH = C2H3 + H2O

and C2H4 + O = CH2CHO + H is also sensitive when the temperature becomes higher.)

Both the sensitivity analysis and reaction pathways analysis reinforce and prove

that the accurate rate coefficient of a concerted elimination reaction is proven to be much

more important than other reactions. The basic chemistry of ethylene may influence overall

reactivity under high-temperatures, so relative rate coefficients should be treated carefully;

but levulinic acid chemistry should be re-examined since it is not sensitive to fuel

consumption in the current mechanism.


91

0.006
EL 0.08
0.005 C2H4
CH4
0.004 H2 0.06
Mole fraction

Mole fraction
0.003
O2
0.04 CO2
0.002 CO

0.02
0.001

0.000 0.00

600 700 800 900 1000 600 700 800 900 1000
Temperature (K) Temperature (K)
0.006
EL
C2H4 0.04
0.005
CH4
H2
0.004 0.03

O2
Mole fraction

Mole fraction
0.003
0.02 CO2
CO
0.002

0.01
0.001

0.000 0.00

600 700 800 900 1000 600 700 800 900 1000
Temperature (K) Temperature (K)

0.005 0.020

0.004
0.015
EL
CO2
Mole fraction

Mole fraction

0.003 C2H4
CH4 0.010 O2
0.002 H2 CO

0.005
0.001

0.000
0.000

600 700 800 900 1000 600 700 800 900 1000
Temperature (K) Temperature (K)

Figure 5.10: Updated mechanism of EL oxidation with experimental data identical to Fig

5.4. Solid lines represent updated model. Dashed lines are from previous simulation.
92

+ +

45% 100%
13% 25% 100%
53% .
.
6% 6%
5.1% 5.1% + R•
+ R•

80%
84%

+
10%
8.9% 50% 28%
52% + R• 22.4%

100% 85% .
100% 81%
100%
100%

Figure 5.11: Reaction path analysis for EL/O2 mixture at 1atm, 800 K (red) and 1000 K

(black), φ = 1. Percentages represent decomposition percent from parent to daughter

species.
93

Ethyl levulinate, φ = 1, P = 1 bar, T = 800 K

2CH3(+M)<=>C2H6(+M)
CH3+HO2<=>CH4+O2
H2O2+O2<=>2HO2
CH2O+HO2<=>HCO+H2O2
CH3O2+CH3<=>2CH3O
CH3O2+CH2O<=>CH3O2H+HCO
EL+HO2<=>EL4J+H2O2
EL+HO2<=>EL3J+H2O2
EL+CH3O2<=>EL4J+CH3O2H
EL+CH3O2<=>EL3J+CH3O2H
CH3+HO2<=>CH3O+OH
H2O2(+M)<=>2OH(+M)
EL<=>C5H8O3+C2H4
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1
Sensitivity Coefficient

Ethyl levulinate, φ = 1, P = 1 bar, T 1000 K


H+O2(+M)<=>HO2(+M)
CH3+HO2<=>CH4+O2
HO2+OH<=>H2O+O2
CH3+OH<=>CH2(S)+H2O
C2H3+O2<=>CH2CHO+O
H2O2(+M)<=>2OH(+M)
C2H4+O<=>CH2CHO+H
C2H4+OH<=>C2H3+H2O
HCO+M<=>H+CO+M
CO+OH<=>CO2+H
CH3+HO2<=>CH3O+OH
H+O2<=>O+OH
EL<=>C5H8O3+C2H4
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4
Sensitivity Coefficient

Figure 5.12: Sensitivity coefficients for 0.5 % ethyl levulinate in N2 under (a) 800 and (b)

1000 K at 1 bar, φ = 1.
94

Chapter 6 CONCLUSION AND FUTURE WORK

The goal of this work was to achieve the following objectives: To build and validate

a new jet-stirred reactor apparatus to study ethyl levulinate (EL) oxidation; iso-octane,

cyclohexene, and n-heptane were used as reference fuels for testing and validation.

Experimental results were also used to validate a chemical kinetic model for a better

understanding of fuel chemistry. Some small intermediates were observed and quantified

by gas chromatography, combined with a thermal conductivity detector, a mass

spectrometer detector and a flame ionization detector. Several species were detected and

quantified: ethylene (C2H4), carbon monoxide (CO), carbon dioxide (CO2), propylene

(C3H6), methane (CH4), hydrogen (H2), oxygen (O2) and 1,3-butadiene (C4H6). Some

oxygenated species were detected, but are not yet quantified: methyl vinyl ketone (C4H6O),

levulinic acid (C5H8O3), ethyl acrylate (C5H8O2), angelica lactone (C5H6O2) and ethanol

(C2H5OH).

The JSR setup was adjusted several times, based on experimental data and

compared to results from Zhang et al. [100]. Different sampling methods were incorporated:

first without a sampling probe, with the addition of a sampling tube, and finally, adding a

sample probe with an installed pumping system. It is indicative of every testing result that

the sampling probe with the pumping system were effectively used to avoid further

reactions occurring in the sampling probe. Cyclohexene pyrolysis validated the necessity

of a pumping system at downstream.

Preliminary simulation of EL oxidation was carried out using the original kinetic

model provided by Dr. Dooley, with the concerted decomposition reaction rate coefficient

of ethyl propanoate (1.6×1013exp(-50,000/RT)) [27]. An EL oxidation experiment was


95

conducted next; with the validated apparatus the fuel consumption profile was a good

match with the simulation results. Primary quantified intermediates were ethylene, CO,

and CO2. Other possible oxygenated products include levulinic acid, ethyl acrylate and

methyl vinyl ketone, which were observed via GC/MS/TCD/FID. However, ethylene did

not fit well with the simulation results under the intermediate temperature regime. To

correct the discrepancies between model prediction and experimental results, various rate

coefficients were replaced and updated by applying analogies to different compounds,

including another concerted reaction rate coefficient from the ethylene time history

measurement (3×1014 exp(54,574/RT)) [45]. The experimental data achieved good agreement

with the adjusted model, as ethylene production is twice as fast as the original model from

750 to 850 K. The concerted elimination reaction was proven to be the key reaction, over

all others, by analyzing reaction pathways and sensitivity.

Future work,

1. The primary objective is quantifying oxygenated species, which can be

accomplished by introducing other GC columns, such as DB-1 or DB-5, and by

changing temperature profiles and oven temperatures. A methanizer for

measuring formaldehyde CH2O is desirable.

2. It is recommended that some rate coefficients in the kinetic model be replaced,

since a few are still assumed to be 1.0×1013. Also, some reactions relative to

primary species, such as ethyl acrylate, should be considered, using an analogy

to ethyl propanoate.

3. EL oxidation under atmospheric pressure was conducted in this study; it is

recommended that EL be tested in other devices at different pressure ranges,


96

such as a shock tube or a rapid compression machine (RCM). Design and

construction of a high-pressure JSR device is required, by doing so, the

mechanism can be testified again and the pressure dependence for some critical

rate coefficients understood.

4. Low-temperature chemistry of EL is currently excluded in the model; a group

from Trinity College Dublin will work on the calculation of rate coefficients for

O2 addition.
97

REFERENCES

1. P. Friedlingstein, R.M. Andrew, J. Rogelj, G. P. Peters, J.G. Canadell, R. Knutti,


G. Luderer, M.R. Raupach, M. Schaeffer, D.P. van Vuuren, C. Le Quéré,
Persistent growth of CO2 emissions and implications for reaching climate
targets. Nat. Geosci., 2014. 7: p. 709-715.

2. A. Demirbas, Importance of biodiesel as transportation fuel. Energy Policy, 2007.


35(9): p. 4661-4670.

3. J.J. Bozell, Chemistry connecting biomass and petroleum processing with a


chemical bridge. Science, 2010. 329(5991): p. 522-523.

4. E. Christensen, A.Williams, S. Paul, S. Burton, R. L. McCormick, Properties and


Performance of Levulinate Esters as Diesel Blend Components. Energy Fuels,
2011. 25(11): p. 5422-5428.

5. D. Matras, J.Villermaux, Un réacteur continu parfaitement agité par jets gazeux


pour l'étude cinétique de réactions chimiques rapides. Chem. Eng. Sci., 1973.
28(1): p. 129-137.

6. D. Gielen, F. Boshell, D. Saygin, Climate and energy challenges for materials


science. Nat. Mater., 2016. 15(2): p. 117-120.

7. David A. Guerrieri, P.J.Caffrey, V. Rao, Investigation into the Vehicle Exhaust


Emissions of High Percentage Ethanol Blends. SAE Technical Paper, 1995
(950777).

8. S. Kim, B.E. Dale, Environmental aspects of ethanol derived from no-tilled corn
grain: nonrenewable energy consumption and greenhouse gas emissions.
Biomass Bioenergy, 2005. 28(5): p. 475-489.

9. A.K. Hossain, P.A. Davies, Plant oils as fuels for compression ignition engines: A
technical review and life-cycle analysis. Renew. Energy, 2010. 35: p. 1-10.

10. A.P. Chalkley, Diesel engines for land and marine work. 1912.

11. A.S. Ramadhas, S. Jayaraj, C. Muraleedharan, Use of vegetable oils as I.C. engine
fuels—A review. Renew. Energy, 2004. 29: p. 727-742.

12. M. Naik, L.C. Meher, S.N. Naik, L.M. Das, Production of biodiesel from high
free fatty acid Karanja (Pongamia pinnata) oil. Biomass Bioenergy, 2008. 32: p.
354-357.
98

13. H. Raheman, A.G.Phadatare, Diesel engine emissions and performance from


blends of karanja methyl ester and diesel. Biomass Bioenergy, 2004. 27: p. 393-
397.

14. J.B. Heywood, Internal combustion engine fundamentals. McGraw-Hill, 1988.

15. I. Hunwartzen, Modification of CFR test engine unit to determine octane numbers
of pure alcohols and gasoline-alcohol blends. SAE technical paper, 1982: p. 1-6.

16. W.-D. Hsieh, R.-H.Chen, T.-L. Wu, T.-H. Lin, Engine performance and pollutant
emission of an SI engine using ethanol–gasoline blended fuels. Atmos. Environ.,
2002. 36: p. 403-410.

17. A.K. Agarwal, Biofuels (alcohols and biodiesel) applications as fuels for internal
combustion engines. Prog. Energy Combust. Sci., 2007. 33: p. 233-271.

18. World fuel ethanol production. 2015. Retrived from


http://ethanolrfa.org/resources/industry/statistics/#1454098996479-8715d404-e546

19. T. Gomiero, M.G. Paoletti, Biofuels: Efficiency, Ethics, and Limits to Human
Appropriation of Ecosystem Services. J. Agric. Environ. Ethics, 2010. 23: p. 403-
434.

20. F. Ma, M.A. Hanna, Biodiesel production: a review. Bioresour. Technol., 1999.
70(1): p. 1-15.

21. H.J. Harwood, Oleochemicals as a fuel: Mechanical and economic feasibility. J.


Am. Oil Chem. Soc., 1984. 61(2): p. 315-324.

22. D.Y.Z. Chang, J.H. Van Gerpen, I. Lee, L.A. Johnson, et al., Fuel Properties and
Emissions of Soybean Oil Esters as Diesel Fuel. J. Am. Oil Chem. Soc., 1996.
73(11): p. 1549-1555.

23. Gas Chromatography, Sheffield Hallam University. Retrived from


http://teaching.shu.ac.uk/hwb/chemistry/tutorials/chrom/gaschrm.htm.

24. J.V. Gerpen, Biodiesel processing and production. Fuel Process. Technol., 2005.
86: p. 1097-1107.

25. H.A. Dabbagh, F.Ghobadi, M.R. Ehsani, M. Moradmand, The influence of ester
additives on the properties of gasoline. Fuel, 2013. 104: p. 216-223.

26. H. O'Neal, S. Benson, A Method for Estimating the Arrhenius A Factors for Four-
and Six-Center Unimolecular Reactions. J. Phys. Chem., 1967. 71(9): p. 2903-
2921.
99

27. W.K. Metcalfe, C.T., P. Dagaut, H.J. Curran, J.M. Simmie, A jet-stirred reactor
and kinetic modeling study of ethyl propanoate oxidation. Combust. Flame, 2009.
156(1): p. 250-260.

28. S. Dooley, H.J. Curran, J.M. Simmie, Autoignition measurements and a validated
kinetic model for the biodiesel surrogate, methyl butanoate. Combust. Flame
2008. 153(1-2): p. 2-32.

29. A. M. El-Nahas, M.V. Navarro, J.M. Simmie, J. W. Bozzelli, H.J. Curran, S.


Dooley, W. Metcalfe, Enthalpies of Formation, Bond Dissociation Energies and
Reaction Paths for the Decomposition of Model Biofuels:  Ethyl Propanoate and
Methyl Butanoate. J. Phys. Chem. A, 2007. 111(19): p. 3727-3739.

30. Renewable Fuel Standard program. 2007. U.S. Environmental Protection Agency

31. M. Lewin, E.M. Pearce, Handbook of Fiber Chemistry. 1998.

32. B. Kamm, P.R. Gruber, M. Kamm, D. J. Hayes, S. Fitzpatrick, M.H.B. Hayes,


J.R.H. Ross, Chapter 7, Biorefineries-Industrial Processes and Products: Status
Quo and Future Directions. 2008.

33. B. Girisuta, L.P.B.M. Janssen, H.J. Heeres, Green Chemicals: A Kinetic Study on
the Conversion of Glucose to Levulinic Acid. Chem. Eng. Res. Des., 2006. 84(5):
p. 339-349.

34. J. J. Bozell, L. Moens, D.C. Elliott, Y. Wang, G.G. Neuenscwander, S.W.


Fitzpatrick, R.J. Bilski, J.L. Jarnefeld, Production of levulinic acid and use as a
platform chemical for derived products. Resour. Conserv. Recy., 2000. 28(3-4): p.
227-239.

35. Renewable fuel standard. 2015. Retrived from


http://ethanolrfa.org/resources/industry/statistics/#1454098996479-8715d404-e546

36. F.M.A. Geilen, B. Engendahl, A. Harwardt, W. Marquardt, J. Klankermayer, W.


Leitner, Selective and Flexible Transformation of Biomass-Derived Platform
Chemicals by a Multifunctional Catalytic System. angewandte Chemie, 2010.
122(32): p. 5642-5646.

37. K. Moshammera, S. Vranckx, H.K. Chakravarty, P. Parab, R. X. Fernandes, K.


Kohse-Höinghaus, An experimental and kinetic modeling study of 2-
methyltetrahydrofuran flames. Combust. Flame, 2013. 160(12): p. 2729-2743.

38. H. Joshi, B.R. Moser, J. Toler, W. F. Smith, T. Walker, Ethyl levulinate: A


potential bio-based diluent for biodiesel which improves cold flow properties.
Biomass Bioenergy, 2011. 35(7): p. 3262-3266.
100

39. D. Klemm, B. Heublein, H.-P. Fink, A. Bohn, Cellulose: Fascinating Biopolymer


and sustainable raw material. Angewandte Chemie, 2005. 44(22): p. 3358-3393.

40. C. Sheng, J.L.T. Azevedo, Estimating the higher heating value of biomass fuels
from basic analysis data. Biomass Bioenergy, 2005. 28(5): p. 499-507.

41. D.L. Kaplan, Biopolymers from renewable resources. 1998.

42. T. Lei, Z. Wang, X. Chang, L. Lin, X. Yan, Y. Sun, X. Shi, X. He, J. Zhu,
Performance and emission characteristics of a diesel engine running on optimized
ethyl levulinateebiodieselediesel blends. Energy, 2016. 95: p. 29-40.

43. Center, H.A.R., Lower and Higher Heating Values of Hydrogen and Other Fuels.
U.S. Department of Energy, Energy Efficiency & Renewable Energy, 2015.

44. S. Thion, A.M. Zaras, M. Szori, P. Dagaut, Theoretical kinetic study for methyl
levulinate: oxidation by OH and CH3 radicals and further unimolecular
decomposition pathways. Phys. Chem. Chem. Phys., 2015. 17: p. 23384-23391.

45. M. AlAbbad, B.R. Giri, M. Szőri, A. Farooq, On the high-temperature


unimolecular decomposition of ethyl levulinate. Proc. Combust. Inst., 2017. 36(1):
p. 187-193.

46. P. Dagaut, M.Cathonnet, J.P. Rouan, R. Foulatier, A. Quilgars, J.C. Boettner, F.


Gaillard, H. James, A jet-stirred reactor for kinetic studies of homogeneous gas-
phase reactions at pressures up to ten atmospheres (≈1 MPa). Journal of Physics
E: Scientific Instruments, 1986. 19(3).

47. A. Chakir, M. Belumam, J. C. Boettner, M. Cathonnet, Kinetic Study of N-


Pentane Oxidation. Combust. Sci. Technol., 1991. 77(4-6): p. 239-260.

48. P. Dagaut, M. Reuillon, M. Cathonnet, High Pressure Oxidation of Liquid Fuels


From Low to High Temperature. 1. n-Heptane and iso-Octane. Combust. Sci.
Technol., 1993. 95(1-6): p. 233-260.

49. Cleaner Combustion Developing Detailed Chemical Kinetic Models. 2013:


Springer London.

50. O. Herbinet, B. Husson, Z. Serinyel, M. Cord, V. Warth, R. Fournet, P.-A.


Glaude, B. Sirjean, F. Battin-Leclerc, Z. Wang, M. Xie, Z. Cheng, F. Qi,
Experimental and modeling investigation of the low-temperature oxidation of n-
heptane. Combust. Flame, 2012. 159(12): p. 3455–3471.

51. S. Singh, W.Grosshandler, P.C. Malte, R.W. Crain Jr., Oxides of nitrogen formed
in high-intensity methanol combustion. Symposium (International) on
Combustion, 1979. 17(1): p. 689-699.
101

52. U. Burke, W.K. Metcalfe, S.M. Burke, K.A. Heufer, P. Dagaut, A detailed
chemical kinetic modeling, ignition delay time and jet-stirred reactor study of
methanol oxidation. Combust. Flame, 2016. 165: p. 125-136.

53. C. Togbé, A. Mzé-Ahmed, P. Dagaut, Experimental and Modeling Study of the


Kinetics of Oxidation of Methanol−Gasoline Surrogate Mixtures (M85
Surrogate) in a Jet-Stirred Reactor. Energy Fuels, 2009. 23(4): p. 1936-1941.

54. P. Dagaut, J.C. Boettner, M. Cathonnet, Kinetic modeling of ethanol pyrolysis and
combustion. J. Chim. Phys., 1992. 89(4): p. 867-884.

55. P. Dagaut, C. Togbé, Oxidation kinetics of butanol–gasoline surrogate mixtures


in a jet-stirred reactor: Experimental and modeling study. Fuel, 2008. 87(15-16):
p. 3313-3321.

56. B. Galmiche, C. Togbé, P. Dagaut, F. Halter, and F. Foucher, Experimental and


Detailed Kinetic Modeling Study of the Oxidation of 1-Propanol in a Pressurized
Jet-Stirred Reactor (JSR) and a Combustion Bomb. Energy Fuels, 2011. 25(5): p.
2013-2021.

57. P. Dagaut, S.M. Sarathy, M.J. Thomson, A chemical kinetic study of n-butanol
oxidation at elevated pressure in a jet stirred reactor. Proc. Combust. Inst., 2009.
32(1): p. 229-237.

58. S.M. Sarathy, M.J. Thomson, C. Togbé, P. Dagaut, F. Halter, C. Mounaim-


Rousselle, An experimental and kinetic modeling study of n-butanol combustion.
Combust. Flame, 2009. 156(4): p. 852-864.

59. P. Dagaut, C. Togbé, Experimental and Modeling Study of the Kinetics of


Oxidation of Butanol−n-Heptane Mixtures in a Jet-stirred Reactor. Energy Fuels,
2009. 23(7): p. 3527-3535.

60. C. Togbé, A. Mzé-Ahmed, P. Dagaut, Kinetics of Oxidation of 2-Butanol and


Isobutanol in a Jet-Stirred Reactor: Experimental Study and Modeling
Investigation. Energy Fuels, 2010. 24(9): p. 5244-5256.

61. Z. Serinyel, C. Togbé, G. Dayma, P. Dagaut, An experimental and modeling study


of 2-methyl-1-butanol oxidation in a jet-stirred reactor. Combust. Flame, 2014.
161(12): p. 3003-3013.

62. C. Togbé, P. Dagaut, A. Mzé-Ahmed, P. Diévart, F. Halter, F. Foucher,


Experimental and Detailed Kinetic Modeling Study of 1-Hexanol Oxidation in a
Pressurized Jet-Stirred Reactor and a Combustion Bomb. Energy Fuels, 2010.
24(11): p. 5859-5875.
102

63. G. Dayma, C. Togbé, P. Dagaut, Experimental and Detailed Kinetic Modeling


Study of Isoamyl Alcohol (Isopentanol) Oxidation in a Jet-Stirred Reactor at
Elevated Pressure. Energy Fuels, 2011. 25(11): p. 4986-4998.

64. C. Togbé, F. Halter, F. Foucher, C. Mounaim-Rousselle, P. Dagaut, Experimental


and detailed kinetic modeling study of 1-pentanol oxidation in a JSR and
combustion in a bomb. Proc. Combust. Inst., 2011. 33(1): p. 367-374.

65. A. Mzé-Ahmed, K. Hadj-A., P. Diévart, P. Dagaut, Kinetics of oxidation of a


reformulated jet fuel (1-hexanol/jet A-1) in a jet-stirred reactor: experimental and
modeling study. Combust. Sci. Technol, 2012. 184(7-8): p. 1039-1050.

66. L. Cai, Y. Uygun, C. Togbé, H. Pitsch, H. Olivier, P. Dagaut, S.M. Sarathy, An


experimental and modeling study of n-octanol combustion. Proc. Combust. Inst.,
2015. 35(1): p. 419-427.

67. S. Gaïl, M.J. Thomson, S.M. Sarathy, S.A. Syed, P. Dagaut, P. Diévart, A.J.
Marchese, F.L. Dryer, A wide-ranging kinetic modeling study of methyl butanoate
combustion. Proc. Combust. Inst., 2007. 31(1): p. 305-311.

68. S.M. Sarathy, S. Gaïl, S.A. Syed, M.J. Thomson, P. Dagaut, A comparison of
saturated and unsaturated C4 fatty acid methyl esters in an opposed flow
diffusion flame and a jet stirred reactor. Proc. Combust. Inst., 2007. 31(1): p.
1015-1022.

69. S. Gaïl, S.M. Sarathy, M.J. Thomson, P. Diévart, P. Dagaut, Experimental and
chemical kinetic modeling study of small methyl esters oxidation: Methyl (E)-2-
butenoate and methyl butanoate. Combust. Flame, 2008. 155(4): p. 635-650.

70. M. H. Hakka, H. Bennadji, J. Biet, M. Yahyaoui, B. Sirjean, V. Warth, L.


Coniglio, O. Herbinet, P.A. Glaude, F. Billaud, F. Battin-Leclerc, Oxidation of
methyl and ethyl butanoates. Int. J. Chem. Kinet., 2010. 42(4): p. 226-252.

71. G. Dayma, S. Gaïl, P. Dagaut, Experimental and kinetic modeling study of the
oxidation of methyl hexanoate. Energy Fuels, 2008. 22(3): p. 1469-1479.

72. G. Dayma, C. Togbé, P. Dagaut, Detailed Kinetic Mechanism for the Oxidation of
Vegetable Oil Methyl Esters: New Evidence from Methyl Heptanoate. Energy
Fuels, 2009. 23(9): p. 4254-4268.

73. G. Dayma, S.M. Sarathy, C. Togbé, C. Yeung, M.J. Thomson, P. Dagaut,


Experimental and kinetic modeling of methyl octanoate oxidation in an opposed-
flow diffusion flame and a jet-stirred reactor. Proc. Combust. Inst., 2011. 33(1):
p. 1037-1043.
103

74. P.-A. Glaude, O. Herbinet, S. Bax, J. Biet, V. Warth, F. Battin-Leclerc, Modeling


of the oxidation of methyl esters—Validation for methyl hexanoate, methyl
heptanoate, and methyl decanoate in a jet-stirred reactor. Combust. Flame, 2010.
157(11): p. 2035-2050.

75. M.H. Hakka, P.-A.Glaude, O. Herbinet, F. Battin-Leclerc, Experimental study of


the oxidation of large surrogates for diesel and biodiesel fuels. Combust. Flame,
2009. 156(11): p. 2129-2144.

76. S. Bax, M.H. Hakka, P.-A. Glaude, O. Herbinet, F. Battin-Leclerc, Experimental


study of the oxidation of methyl oleate in a jet-stirred reactor. Combust. Flame,
2010. 157(6): p. 1220-1229.

77. P. Dagaut, S. Gaïl, M. Sahasrabudhe, Rapeseed oil methyl ester oxidation over
extended ranges of pressure, temperature, and equivalence ratio: experimental
and modeling kinetic study. Proc. Combust. Inst., 2007. 31(2): p. 2955-2961.

78. P. Dagaut, S. Gaïl, Chemical Kinetic Study of the Effect of a Biofuel Additive on
Jet-A1 Combustion. J. Phys. Chem. A, 2007. 111(19): p. 3992-4000.

79. G. Dayma, F. Halter., F. Foucher, C. Togbé, C. Mounaim-Rousselle, P. Dagaut,


Experimental and Detailed Kinetic Modeling Study of Ethyl Pentanoate (Ethyl
Valerate) Oxidation in a Jet Stirred Reactor and Laminar Burning Velocities in a
Spherical Combustion Chamber. Energy Fuels, 2012. 26(8): p. 4735-4748.

80. H.P. Ramirez, K. Hadj-Ali, P. Diévart, G. Dayma, C. Togbé, G. Moréac, P.


Dagaut, Oxidation of commercial and surrogate bio-Diesel fuels (B30) in a jet-
stirred reactor at elevated pressure: Experimental and modeling kinetic study.
Proc. Combust. Inst., 2011. 33(1): p. 375-382.

81. Z. Serinyel, C. Togbé, G. Dayma, P. Dagaut, Experimental and Modeling Study of


the Oxidation of Two Branched Aldehydes in a Jet-Stirred Reactor: 2-
Methylbutanal and 3-Methylbutanal. Energy Fuels, 2017. 31(3): p. 3206-3218.

82. A. Rodriguez, O. Herbinet, F. Battin-Leclerc, A study of the low-temperature


oxidation of a long chain aldehyde: n-hexanal. Proc. Combust. Inst., 2017. 36(1):
p. 365-372.

83. S. Thion, C. Togbé, G. Dayma, Z. Serinyel, P. Dagaut, Experimental and Detailed


Kinetic Modeling Study of Cyclopentanone Oxidation in a Jet-Stirred Reactor at 1
and 10 atm. Energy Fuels, 2017. 31(3): p. 2144-2155.

84. S. Thion, P. Diévart, P.V. Cauwenberghe, G. Dayma, Z. Serinyel, P. Dagaut, An


experimental study in a jet-stirred reactor and a comprehensive kinetic
mechanism for the oxidation of methyl ethyl ketone. Proc. Combust. Inst., 2017.
36(1): p. 459-467.
104

85. R. Shahla, C. Togbé, S. Thion, R. Timothée, M. Lailliau, F. Halter, C. Chauveau,


G. Dayma, P. Dagaut, Burning velocities and jet-stirred reactor oxidation of
diethyl carbonate. Proc. Combust. Inst., 2017. 36(1): p. 553-560.

86. H.J. Curran, W.J. Pitz, C.K. Westbrook, P. Dagaut, J.-C. Boettner, M. Cathonnet,
A Wide Range ModelingStudy of Dimethyl Ether Oxidation. Int. J. Chem. Kinet.,
1998. 30(3): p. 229-241.

87. C.A. Daly, J.M. Simmie, P. Dagaut, M. Cathonnet, Oxidation of


dimethoxymethane in a jet-stirred reactor. Combust. Flame, 2001. 125(3): p.
1106-1117.

88. P.A. Glaude, F. Battin-Leclerc, B. Judenherc, V. Warth, R. Fournet, G.M. Côme,


G. Scacchi, P. Dagaut, M. Cathonnet, Experimental and modeling study of the
gas-phase oxidation of methyl and ethyl tertiary butyl ethers. Combust. Flame,
2000. 121(1-2): p. 345-355.

89. W.W. Ayass, E.F. Nasir, A. Farooq, S.M. Sarathy, Mixing-structure relationship
in jet-stirred reactors. Chem. Eng. Res. Des., 2016. 111: p. 461-464.

90. P. G. Lignola, E. Reverchon, A Jet Stirred Reactor for Combustion Studies:


Design and Characterization. Combust. Sci. Technol., 1988. 60(4-6): p. 319-333.

91. D.R. Matras, Rules for the construction and extrapolation of reactors self-stirred
by gas jets. Can. J. Chem. Eng., 1975. 53: p. 297-300.

92. O. Herbinet, P.-M. Marquaire, F. Battin-Leclerc, R. Fournet, Thermal


decomposition of n-dodecane: Experiments and kinetic modeling. J. Anal. Appl.
Pyrolysis, 2007. 78(2): p. 419-429.

93. F. Battin-Leclerc, O. Herbinet, P.-A. Glaude, R. Fournet, Z. Zhou, L. Deng, H.


Guo, M. Xie, F. Qi, New experimental evidences about the formation and
consumption of ketohydroperoxides. Proc. Combust. Inst., 2011. 33(1): p. 325-
331.

94. O. Herbinet, F. Battin-Leclerc, S. Bax, H.L. Gall, P.-A. Glaude, R. Fournet, Z.


Zhou, L. Deng, H. Guo, M. Xie, F. Qi, Detailed product analysis during the low
temperature oxidation of n-butane. Phys. Chem. Chem. Phys., 2011. 13: p. 296-
308.

95. Z. Wang, L. Zhang, K. Moshammer, D.M. Popolan-Vaida, V.S.B. Shankar, A.


Lucassen, C. Hemken, C.A. Taatjes, S.R. Leone, K. Kohse-Höinghaus, N.
Hansen, P. Dagaut, S.M. Sarathy, Additional chain-branching pathways in the
low-temperature oxidation of branched alkanes. Combust. Flame, 2016. 164: p.
386-396.
105

96. R.L. Grob, E.F. Barry, Modern Practice of Gas Chromatography. 2004.

97. C. Wang, Parallel GC for Complete Refinery Gas Analysis. 2007.

98. W.K. Cheng, T. Summers, N. Collings, The fast-response flame ionization


detector. Prog. Energy Combust. Sci., 1998. 24(2): p. 89-124.

99. J.T. Scanlon, D.E. Willis, Calculation of Flame Ionization Detector Relative
Response Factors Using the Effective Carbon Number Concept. J. Chromatogr.
Sci, 1985. 23(8): p. 333-340.

100. K. Zhang, C. Banyon, J. Bugler, H.J. Curran, A. Rodriguez, O. Herbinet, F.


Battin-Leclerc, C. B'Chir, K.A. Heufer, An updated experimental and kinetic
modeling study of n-heptane oxidation. Combust. Flame, 2016. 172: p. 116-135.

101. W. Tsang, C.M. Rosado-Reyes, Unimolecular Rate Expression for Cyclohexene


Decomposition and Its Use in Chemical Thermometry under Shock Tube
Conditions. J. Phys. Chem. A, 2015. 119(28): p. 7155-7162.

102. N. Atef, G. Kukkadapu, S.Y. Mohamed, M.A. Rashidi, C. Banyon, M. Mehl,


K.A. Heufer, E.F. Nasir, A. Alfazazi, A.K. Das, C.K. Westbrook, W.J. Pitz, T.
Lu, A. Farooq, C.-J. Sung, H.J. Curran, S.M. Sarathy, A comprehensive iso-
octane combustion model with improved thermochemistry and chemical kinetics.
Combust. Flame, 2017. 178: p. 111-134.

103. A. Chakir, M. Bellimam, J.C. Boettner, M. Cathonnet, Kinetic study of n-heptane


oxidation. Int. J. Chem. Kinet., 1992. 24(4): p. 385-410.

104. H.M. Hakka, R.F.C., A. Pekalski, P.-A. Glaude, F. Battin-Leclerc, Experimental


and modeling study of ultra-rich oxidation of n-heptane. Fuel, 2015. 144: p. 358-
368.

105. Philippe Dagaut, M.R., Michel Cathonnet, Experimental study of the oxidation of
n-heptane in a jet stirred reactor from low to high temperature and pressures up
to 40 atm. Combust. Flame, 1995. 101(1-2): p. 132-140.

106. Jiaxiang Zhang, S.N., Yingjia Zhang, Chenglong Tang, Xue Jiang, Erjiang Hu,
Zuohua Huang, Experimental and modeling study of the auto-ignition of n-
heptane/n-butanol mixtures. Combust. Flame, 2013. 160(1): p. 31-39.

107. D.F. Davidson, Z.H., G.L. Pilla, A. Farooq, R.D. Cook, R.K. Hanson, Multi-
species time-history measurements during n-heptane oxidation behind reflected
shock waves. Combust. Flame, 2010. 157(10): p. 1899-1905.
106

108. J. Herzler, L.J., P. Roth, Shock tube study of the ignition of lean n-heptane/air
mixtures at intermediate temperatures and high pressures. Proc. Combust. Inst.,
2005. 30(1): p. 1147-1153.

109. H.K. Ciezki, G. Adomeit, Shock-tube investigation of self-ignition of n-heptane-


air mixtures under engine relevant conditions. Combust. Flame, 1993. 93(4): p.
421-433.

110. C.M. Coats, A.Williams, Investigation of the ignition and combustion of n-


heptane-oxygen mixtures. Symposium (International) on Combustion, 1979.
17(1): p. 611-621.

111. H.J. Curran, P. Gaffuri, W.J. Pitz, C.K. Westbrook, A Comprehensive Modeling
Study of n-Heptane Oxidation. Combust. Flame, 1998. 114(1-2): p. 149-177.

112. K. Zhang, C. Banyon, J. Bugler, H.J. Curran, A. Rodriguez, O. Herbinet, F.


Battin-Leclerc, C. B'Chir, K.A. Heufer, An updated experimental and kinetic
modeling study of n-heptane oxidation. Combust. Flame, 2016. 172: p. 116-135.

113. O. Herbinet, B. Husson, Z. Serinyel, M. Cord, V. Warth, R. Fournet, P.-A.


Glaude, B. Sirjean, F. Battin-Leclerc, Z. Wang, M. Xie, Z. Cheng, F. Qi,
Experimental and modeling investigation of the low-temperature oxidation of n-
heptane. Combust. Flame, 2012. 159(12): p. 3455-3471.

114. CHEMKIN-PRO. Reaction Design, 2013. 15131.

115. M. Uchiyama, A. Amano, T. Tomioka, Thermal Decomposition of Cyclohexene.


J. Phys. Chem., 1964. 68: p. 1878-1881.

116. P.G. Lignola, F.P.D. Maio, A. Marzocchella, R. Mercogliano, E. Reverchon,


JSFR combustion processes of n-heptane and isooctane. Symposium
(International) on Combustion, 1989. 22(1): p. 1625-1633.

117. A. D'Anna, R. Mercoguano, R. Barrbella, A. Ciajolo, Low Temperature Oxidation


Chemistry of iso-Octane under High Pressure Conditions. Combust. Sci.
Technol., 1992. 83: p. 217-232.

118. Z. Serinyel, G. Black, H.J. Curran, J.M. Simmie, A Shock Tube and Chemical
Kinetic Modeling Study of Methy Ethyl Ketone Oxidation. Combust. Sci.
Technol., 2010. 182(4-6): p. 574-587.

119. S. Thion, P. Diévart, P.V. Cauwenberghe, G. Dayma, Z. Serinyel, An


experimental study in a jet-stirred reactor and a comprehensive kinetic
mechanism for the oxidation of methyl ethyl ketone. Proc. Combust. Inst., 2017.
36(1): p. 459-467.
107

120. M. Mehl, W.J. Pitz, C.K. Westbrook, H.J. Curran, Kinetic modeling of gasoline
surrogate components and mixtures under engine conditions. Proc. Combust.
Inst., 2011. 33(1): p. 193-200.

121. J.M. Hudzik, J.W. Bozzelli, Thermochemistry and Bond Dissociation Energies of
Ketones. J. Phys. Chem. A, 2012. 116(23): p. 5707–5722.

122. A.M. El-Nahas, M.V. Navarro, J.M. Simmie, J.W. Bozzelli, H.J. Curran, S.
Dooley, W. Metcalfe, Enthalpies of Formation, Bond Dissociation Energies and
Reaction Paths for the Decomposition of Model Biofuels:  Ethyl Propanoate and
Methyl Butanoate. J. Phys. Chem. A, 2007. 111(19): p. 3727–3739.

123. Y. Zhang, H.J. Curran, G. Capriolo, S. Dooley, K. Djebbi, S.M. Sarathy, A.


Farooq, Ignition delays and a combustion kinetic model of the lignocellulosic
biofuel, ethyl levulinate. Personal Communication.

124. F. Dryer, D. Naegeli, I. Glassman, Temperature dependence of the reaction CO +


OH = CO2 + H. Combust. Flame, 1971. 17(2): p. 270-272.

125. H.J. Curran, P.Gaffuri, W.J. Pitz, C.K. Westbrook, A comprehensive modeling
study of iso-octane oxidation. Combust. Flame, 2002. 129(3): p. 253-280.

126. A.T. Blades, P.W. Gilderson, Kinetics of the Thermal Decomposition of Ethyl
Propionate. Can. J. Chem., 1960. 38(9): p. 1412-1415.

127. W.K. Metcalfe, S. Dooley, H.J. Curran, J.M. Simmie , A.M. El-Nahas, M.V.
Navarro, Experimental and Modeling Study of C5H10O2 Ethyl and Methyl Esters.
J. Phys. Chem. A, 2007. 111(19): p. 4001-4014.

128. C.K. Westbrook, W.J. Pitz, P.R. Westmoreland, F.L. Dryer, M. Chaos, P.
Osswald, K. Kohse-Höinghaus, T.A. Cool, J. Wang, B. Yang, N. Hansen, T.
Kasper, A detailed chemical kinetic reaction mechanism for oxidation of four
small alkyl esters in laminar premixed flames. Proce. Combust. Inst., 2009. 32(1):
p. 221-228.

129. J.P. Orme , H.J. Curran, J.M. Simmie, Experimental and Modeling Study of
Methyl Cyclohexane Pyrolysis and Oxidation. J. Phys. Chem. A., 2006. 110(1): p.
114-131.

130. Z. Serinyel, N. Chaumeix, G. Black, J.M. Simmie, H.J. Curran, Experimental and
Chemical Kinetic Modeling Study of 3-Pentanone Oxidation. J. Phys. Chem. A.,
2010. 114(46): p. 12176-12186.

131. M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R.
Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji,
M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L.
108

Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida,


T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E.
Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N.
Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant,
S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.
B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O.
Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K.
Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S.
Dapprich, A.D. Daniels, Ö. Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J.
Fox, 2009.

132. E.R. Ritter, J.W.Bozzelli, THERM: Thermodynamic property estimation for gas
phase radicals and molecules. Int. J. Chem. Kinet., 1991. 23: p. 767-778.

133. Z. Serinyel, G.B., H.J. Curran, J.M. Simmie, A shock tube and chemical kinetic
modeling study of methyl ethyl ketone oxidation. Combust. Sci. Technol., 2010.
182: p. 574-587.

134. Olivier Herbinet, W.J.P., Charles K. Westbrook, Detailed chemical kinetic


oxidation mechanism for a biodiesel surrogate. Combust. Flame, 2008. 154: p.
507-528.

135. C. Franklin Goldsmith, W.H.G., Stephen J. Klippenstein, Role of O2 + QOOH in


Low-Temperature Ignition of Propane. 1. Temperature and Pressure Dependent
Rate Coefficients. J. Phys. Chem. A, 2012. 116(13): p. 3325-3346.

136. J. Zádor, C.A. Taatjes, R.X. Fernandes, Kinetics of elementary reactions in low-
temperautre autoignition chemistry. Proc. Combust. Inst., 2011. 37(4): p. 371-
421.
109

SUPPLEMENTAL MATERIAL

Low-temperature Oxidation Chemistry

In all engines, including spark-ignition, compression-ignition and homogeneous charge

compression ignition engines, fuel reacts with oxygen at low temperatures and gradually

approaches to the high temperature region because of piston compression. Note that some

high CN species, such as long-chain biodiesels, start to be consumed at much lower

temperatures (500 to 600 K), producing small aldehydes and alkenes [134]. Afterwards,

the negative temperature coefficient zone emerges and total reactivity decreases as the

temperature increases. Eventually, fuel is again consumed and ignites rapidly at high

temperature. These pre-ignition phenomena are known as low temperature chemistry.

For some bio-derived fuels with large aliphatic chains attached to a methyl or ethyl

ester, low temperature ignition chemistry has been found to show a similar pattern to large

n-alkanes [134]. This unusual phenomenon is interpreted as negative temperature

coefficient (NTC) [109], where fuel concentration increases with temperature enhancement

in a certain temperature range. At low temperatures (from 500 to 700 K), auto-ignition

occurs after a series of O2 additions, isomerization and chain branching, as described in

Fig. 8.1. Reactions are initiated from fuel abstraction by O2 to form the fuel radicals R•.

These fuel radicals react with O2 again and are converted, primarily to alkylperoxyl radicals

ROO• , which undergo hydrogen atom internal isomerization to hydroperoxyalkyl

radicals •QOOH. The isomerization proceeds via five, six, seven and eight-membered

transition state rings, in which the six-membered ring is the most favorable transition state

[111, 135]. Or, ROO• can decompose to cyclic ethers + hydroxyl radicals •OH by breaking

the O-O single bond or by forming olefins + hydroperoxyl radicals HO2•. Since the free
110

radical site •QOOH is located at the carbon atom, the main pathway for •QOOH is a second

O2 addition, which forms peroxy hydroperoxyl alkyl radicals •OOQOOH. Eventually,

chain branching occurs, mainly by •OOQOOH isomerization, to keto-hydroperoxide

and •OH radicals, the major path for low temperature chemistry (Fig. 8.2 blue region). Next,

exothermic chain branching forms keto-hydroperoxide, which promotes reactivity at low

temperatures and releases more heat that favors the rate of ROO• and •QOOH producing

olefins + hydroperoxyl radicals HO2• and cyclic ethers + hydroxyl radicals •OH.

Shown in the orange region in Fig. 8.2, lower reactivity results in propagations

of •QOOH and ROO• to more HO2• radicals and olefins (relatively more stable species and

radicals than •OH radicals). This competition is responsible for the decreased reactivity

and for the NTC phenomena. The NTC region ends when HO2 builds up and self-reacts to

form H2O2 + O2, where H2O2 forms 2 •OH radicals (chain branching) and increases

reactivity again [135].

At high temperatures (green region in Fig. 8.2), H atom abstraction and unimolecular

fuel decomposition are the main pathways where alkyl radicals R• undergo beta-scission

to smaller olefins and other species. Furthermore, the •H radical, reacting with O2 to

produce •O + •OH, is the dominant chain branching reaction. Two radicals keep the

combustion environment reactive, consuming most fuel molecules and generating smaller

and more stable species like H2O and CO2. Simulation and experimental results of

biodiesels were also compared with long-chain hydrocarbons so that both species can be

observed with this series of reaction classes [134].


111

RH
Initiation
R• β-scission products

O2 addition

RO2•

Isomerization HO2 • + Olefin

• QOOH OH • + cyclic ether

nd
2 O2 addition R’ • + β-scission products

• OOQOOH

Keto-hydroperoxide +OH

Figure 8.1: General reaction scheme of low temperature oxidation of alkanes (RH)

with long connected secondary carbons [136].


112

Figure 8.2: Experimental (plots) [109] and modeling (lines) results for n-heptane/air

mixture ignition delay times under multiple conditions [100]. Graph is divided into green,

orange and blue, governed by chain branching to •O + •OH, chain propagation to HO2• +

olefins and chain branching to ketohydroperoxide + •OH, respectively.

Introduction to Kinetics and Thermodynamics

Thermodynamics

Using the thermodynamic properties: heat capacity 𝐶, entropy S, and enthalpy H,

one can predict and understand the combustion phenomena. As heat transferring to/from

the system, heat capacity 𝐶𝑣 is defined as the internal energy change 𝜕𝑈 by changing per

unit temperature at constant volume:

𝜕𝑈
𝐶𝑣 = ( )
𝜕𝑇 𝑣

For an isobaric system, the heat capacity 𝐶𝑝 is defined as the enthalpy difference by

changing per unit temperature at constant pressure, in the following expression:

𝜕𝐻
𝐶𝑝 = ( )
𝜕𝑇 𝑝

Furthermore, at a defined temperature, the enthalpy of the system can be calculated by

integration:
𝑇2
𝐻(𝑇2 ) = 𝐻(𝑇1 ) + ∫ 𝐶𝑝 𝑑𝑇
𝑇1

Where 𝐻(𝑇1 ) is a standard enthalpy usually at 273 or 298 K for difference substance.

Moreover, when it comes to a closed system, the second law of thermodynamics

define the information about the direction of heat change and determine if the heat transfer
113

is spontaneous or not, as well as reversibility. In a reversible thermodynamic process, when

heat transfer to a system happens, a variable entropy ∆𝑆 based on the expression:

𝑑𝑞𝑟𝑒𝑣
𝑑𝑆 =
𝑇

A reversible pathway between two states 𝑖 and 𝑓 of a system can be integrated:

𝑓
𝑑𝑞𝑟𝑒𝑣
∆𝑆 = ∫
𝑖 𝑇

To meet the criterion of second law of thermodynamics, entropy should always be positive

from initial to final state when a spontaneous change occurs, giving an expression for an

irreversible process:

∆𝑆𝑖𝑟𝑟𝑒𝑣 > 0

Just like enthalpy, this physical quantity can also be integrated for a reversible process:
𝑇𝑓
𝑑𝑞𝑟𝑒𝑣
𝑆(𝑇𝑓 ) = 𝑆(𝑇𝑖 ) + ∫
𝑇𝑖 𝑇

Since total entropy can be interpreted as ∆𝑆𝑖𝑛𝑡 + ∆𝑆𝑒𝑥𝑡 where ∆𝑆𝑖𝑛𝑡 is denoted as the

entropy change in the system and ∆𝑆𝑒𝑥𝑡 represents the entropy change from the

𝑄
surrounding. ∆𝑆𝑒𝑥𝑡 can be replaced by − 𝑇 (Q is the heat transfer to the system). The

irreversible process of entropy change can be arranged as below, also named Clausius

inequality:

𝑄
∆𝑆𝑖𝑛𝑡 ≥
𝑇

As it is rearranged:

0 ≥ 𝑄 − 𝑇∆𝑆

In an isobaric system, the heat transfer 𝑄 is equal to enthalpy change ∆𝐻 that Gibbs energy

is introduced for a convenient expression,


114

∆𝐺 = ∆𝐻 − 𝑇∆𝑆

For better understanding of the spontaneity of a system or a reaction, the conclusion can

be made that a reaction is spontaneous if ∆𝐺 is negative and ∆𝐺 = 0 indicates the criteria

for equilibrium.

Reaction Kinetics

Experimentally, many rate coefficients exhibit a straight line while plotting rate

coefficient in natural log form against 1/T with a negative slope. The corresponding

equation can be written as below, also known as Arrhenius equation, where 𝐴 is pre-

exponential factor, 𝐸𝑎 is activation energy, 𝑅 is the universal gas constant and 𝑇 is

temperature.

𝐸𝑎
ln 𝑘(𝑇) = ln 𝐴 −
𝑅𝑇

The pre-exponential factor, also called frequency factor, is to determine the number of

times two molecules colliding with each other. According to collision theory, the activation

energy can be defined as the minimum energy required to overcome the potential energy

barrier to proceed a reaction. As far as the energy is enough to surpass the barrier, two

molecules form a transition-state complex which is relatively unstable with short lifetime.

Eventually, the transition state molecule breaks into products by releasing energy in terms

of heat and the equation can be organized as an alternative form with separate temperature

coefficient:

−𝐸𝑎
𝑘(𝑇) = 𝐴𝑇 𝑛 exp( )
𝑅𝑇

To solve the reverse rate coefficients for thousands of reactions, the thermodynamic aspects

of chemical kinetics play an important role. Based on the fundamental concept of


115

thermodynamics, Gibbs energy of activation for a reaction can be determined, relative to

an equilibrium constant:

∆𝐺 = −𝑅𝑇 𝑙𝑛𝐾

Since an equilibrium constant is the ratio of forward and backward rate coefficients for a

reaction, the above equation becomes:

∆𝐺
𝑘𝑓 = 𝑘𝑏 exp(− )
𝑅𝑇

𝑘𝑓 is the rate coefficient of a forward reaction and 𝑘𝑏 for backward reaction. By replacing

Gibbs energy with entropy and enthalpy of activation, the equation can be rearranged as:

∆𝑆 ∆𝐻
𝑘𝑓 = 𝑘𝑏 exp( )exp(− )
𝑅 𝑅𝑇

Therefore, combining the chemical kinetics, rate coefficients of backward reactions can be

calculated from two thermodynamic quantities: entropy and enthalpy.

S-ar putea să vă placă și