<<

COMPUTATIONAL STUDIES OF PROCESSES AND OXYGENATED SPECIES

DISSERTATION

Presented in Partial Fulfillment of the Requirements for

the Degree of Doctor of Philosophy

in the Graduate School of The Ohio State University

By

Carrigan Jo Hayes, B.S.

****

The Ohio State University 2007

Dissertation Committee: Approved by:

Dr. Christopher M. Hadad, Advisor ______Dr. Matthew S. Platz Advisor Graduate Program in Chemistry Dr. David J. Hart

i

ABSTRACT

Within this dissertation, we report on explorations of reactive species with

implications for combustion and atmospheric chemistry. Various computational

approaches, including density functional theory (DFT) and master equation methods,

were used to complete these projects.

The majority of this thesis involves the oxidation pathways of the alkylated

heterocycles that provide a model framework for understanding coal combustion. The

enthalpies and energies of reaction for -atom loss and -group

fragmentations at various temperatures were calculated via density functional theory

(B3LYP/6-311+G**//B3LYP/6-31G*); these results were calibrated against CBS-QB3

calculations. It was suggested that both hydrogen-atom loss and alkyl-group loss reactions will contribute as initiation steps for the high-temperature combustion reactions of these rings. Longer alkyl chains will increase reactivity, and the azabenzene units are more likely to react than the five-membered heterocyclic rings. The initial steps of formation are expected to become more favorable at high temperatures.

The oxidation steps of these radicals were shown to be exothermic and exoergic,

as expected. DFT studies (B3LYP/6-311+G**//B3LYP/6-31G*) showed that these

resultant peroxy radicals were more likely to undergo intramolecular reactions to form

bicyclic structures. Furthermore, several pathways seemed feasible and must be

ii considered in rationalizing coal chemistry: formation of either a four-membered or five-

membered ring occurs with similar kinetics and thermodynamics; cyclization at

to form a species has a high reaction barrier but is ultimately quite exoergic. An

internal H-atom transfer can occur on the substituted side chain with a low barrier and

favorable energy of reaction.

We have completed other projects with implications for combustion chemistry

and oxygenated species. 1,5-H transfers proceed readily for n-propylperoxy radical due

to this radical’s ability to adopt a six-membered transition state. This reaction and others

available to this species are of interest due to the ability of n-propylperoxy radical to

serve as a model compound for understanding the combustion of larger

fuels. Conformational possibilities for this species were explored via B3LYP/6-31G*

and mPW1K/6-31+G** levels of theory to ensure that rotational barriers would not compete with energies of reaction. It was seen that rotamer interconversion occurs with barrier heights of less than 5 kcal/mol, far less than the relevant reaction activation barriers (~20-25 kcal/mol); thus, rotamer interconversion was not expected to affect overall energetics. Building on these results, the unimolecular decomposition of propan-

1-ol-1-peroxy radical was similarly modeled using DFT methods. It was seen than the quantitative energetics of the relevant decompositions were very similar to those of the hydrocarbon analogue, although a wider variety of functionalized products were formed.

Complexes of with various solvents were modeled to better understand

certain spectroscopic phenomena and potential atmospheric behaviors of oxygenated

species. Experimental work on these complexes had noted a red shift due to

iii complexation of ethanol with that was not seen with any other solvents.

Theoretical spectra were generated using HF/6-31G* and MP2/6-31G* optimizations and

compared well to the experimental spectra. The red shift seen in benzene was attributed

to an interaction of ethanol with the pi system of the benzene ring.

Finally, engine performance varies given the fuel of interest. Hydrogen has often

been proposed as an alternative fuel that would improve engine performance and

minimize harmful emissions. Its use as a fuel additive was explored with n-heptane and

the primary reference fuels (a mixture of iso-octane and n-heptane), using the master equation program CHEMKIN 4.1. It was seen that hydrogen augmentation did increase flame speed and decrease monoxide emissions.

iv

ACKNOWLEDGMENTS

I would like to thank Dr. Christopher Hadad for his advice, support, and guidance over my time at The Ohio State University; I have learned a great deal both from his words as an advisor and his example as a teacher. I thank members of the Hadad group, past and present, for their thoughtful explanations and thorough discussions. I have worked with Reaction Design and McMaster Fuel and am indebted to these companies’ expertise and support, as well as to research support from the NSF-funded Environmental

Molecular Science Institute and the Ohio Supercomputer Center. I have received a

University Fellowship, a Grilly Scholarship, and two GAANN Fellowships; I am grateful to the Graduate School and the Department of Chemistry for these awards.

I thank my family for their encouragement and love, especially my parents, Steve and Karen Hayes; my siblings, Rob and Katie; and my grandparents, Tom and Jeanne

Wright. Although they are too numerous to mention by name, I also thank many uncles, aunts, cousins, and other relatives whose kind support has been a continued source of inspiration throughout my education. In particular, I mention the relatives whom have passed away during my time in graduate school: my aunt, Judy Boling; my great- grandfather, Fred Hines; my uncle, Gary Wright; and my great-aunt, JoAnn Grigsby. I hope to approach my with the compassion, intellect, humor, and faith that they demonstrated throughout theirs.

v

VITA

September 2, 1980- Born: Findlay, Ohio

May 20, 2002- B.S. in Chemistry: Ohio Northern University (Ada, Ohio)

2002-2003- University Fellow, Graduate School, The Ohio State University

2004-2005- GAANN Fellow, OSU Department of Chemistry

2002-present- Grilly Scholar, OSU Department of Chemistry

2003-present- Graduate Research and Teaching Associate, OSU Department of Chemistry

PUBLICATIONS

1. Zalyubovsky, S. J.; Glover, B. G.; Miller, T. A.; Hayes, C.; Merle, J. K.; Hadad, C. M. “Observation of the A-X Electronic Transition of the 1-C3H7O2 and 2- C3H7O2 Radicals Using Cavity Ringdown .” J. Phys. Chem. A 2005, 109, 1308 - 1315.

2. Merle, J. K.; Hayes, C. J.; Zalyubovsky, S. J.; Glover, B. G.; Miller, T. A.; Hadad, C. M. “Theoretical Determinations of the Ambient Conformational Distribution and Unimolecular Decomposition of n-Propylperoxy Radical.” J. Phys. Chem. A 2005, 109, 3637 - 3646.

3. Levering, L. M.; Hayes, C. J.; Callahan, K. M.; Hadad, C. M.; Allen, H. C. “Non-Aqueous Solvation of n-Octanol and Ethanol: Spectroscopic and Computational Studies.” J. Phys. Chem. B, 2006, 110, 6325-6331.

FIELDS OF STUDY

Major field: Chemistry

vi

TABLE OF CONTENTS PAGE Abstract………………………………………………………………………. ii Acknowledgements…………………………………………………………... v Vita………………………………………………………………………….... vi List of Tables…………………………………………………………………. xi List of Figures……………………………………………………………….... xv

CHAPTERS: 1. Chemistry of reactive radical intermediates of combustion and atmospheric reactions………………………………………………… 1

1.1 Introduction…………………………………………………………... 1 1.2 Basic Concepts of Combustion Chemistry…………………………... 1 1.2.1 Free radicals……………………………………………...... 2 1.2.2 Combustion at varying temperatures………………………... 4 1.2.2.1 High-temperature combustion……………………. 5 1.2.2.2 Low-temperature combustion…………………….. 6 1.2.2.3 Negative temperature coefficient (NTC) range…... 9 1.2.2.4 Atmospheric oxidation……………………………. 11 1.2.3 Methods for studying reactive combustion species…………. 15 1.2.3.1 Experimental methods……………………………. 15 1.2.3.2 Computational methods…………………………... 18 1.3 Reactive Radical Intermediates in Combustion Chemistry………….. 21 1.3.1 Aliphatic systems……………………………………………. 21 1.3.1.1 combustion……………………………... 21 1.3.1.2 Ethyl radical + O2...... 24 1.3.1.3 n-Propylperoxy radical…………………………... 27 1.3.1.4 n-Butoxy radical…………………………………. 32 1.3.1.5 1-Pentyl radical…………………………………... 33 1.3.1.6 Larger aliphatic systems…………………………. 33 1.3.2 Aromatic systems……………………………………………. 35 1.3.2.1 Soot formation…………………………………… 35 1.3.2.2 Benzene and toluene……………………………... 36 1.3.2.3 Benzene oxidation………………………………... 38

vii 1.3.2.4 Alkylated aromatics……………………………… 49 1.3.2.5 Heteroaromatic combustion……………………… 51 1.3.2.6 Alkylated heteroaromatics……………………….. 60 1.3.2.7 Peroxy radicals of alkylated heteroaromatics……. 68 1.4 Future Challenges in Combustion Chemistry………………………... 77 1.4.1 Fuel additives………………………………………………... 78 1.4.2 Biodiesel…………………………………………………….. 81 1.5 Conclusions…………………………………………………………... 84 References………………………………………………………………….. 86

2. Combustion Pathways of the Alkylated Heteroaromatics: Bond Dissociation Enthalpies and Alkyl-Group Fragmentations…………... 109

2.1 Introduction………………………...... 109 2.2 Computational Methods……………………………...... 116 2.3 Results………………………………………………………………. 118 2.4 Discussion…………………………………………...... 118 2.4.1 Toluene as reference compound……………………………. 119 2.4.2 Non-alkylated heteroaromatics……………………………... 120 2.4.2.1 Azabenzenes……………………………...... 120 2.4.2.2 Five-membered rings…………………………….. 120 2.4.3 Methyl-substituted heteroaromatics: H-atom loss………….. 121 2.4.3.1 Azabenzenes……………………………...... 121 2.4.3.2 Five-membered rings…………………………….. 125 2.4.4 Ethyl-substituted heteroaromatics: H-atom loss……………. 126 2.4.4.1 Azabenzenes……………………………...... 127 2.4.4.2 Five-membered rings…………………………….. 127 2.4.5 Methyl-substituted heteroaromatics: methyl loss…………... 127 2.4.5.1 Azabenzenes……………………………...... 128 2.4.5.2 Five-membered rings…………………………….. 129 2.4.6 Ethyl-substituted heteroaromatics: ethyl loss………………. 129 2.4.7. General trends………………………………………………. 130 2.4.7.1 Spin density………………………………………. 130 2.4.7.2 Geometry………………………………………… 133 2.4.7.3 Temperature effects……………………………… 134 2.4.7.4 Approximations………………………………….. 140 2.5 Conclusions………………………………………………………… 142 References…………………………………………………………………. 144

viii 3. Combustion Pathways of the Alkylated Heteroaromatics: Peroxy Radical Pathways of the Alkylated Azabenzenes…………………….. 150

3.1 Introduction…………………………………………………………... 150 3.2 Computational Methods……………………………………………… 156 3.3 Results………………………………………………………………... 158 3.4 Discussion……………………………………………………………. 159 3.4.1 Toluene as reference compound……………………………... 159 3.4.2 Methyl-substituted (picolines)…………………….. 163 3.4.3 Temperature effects for the picolinylperoxy radicals………... 172 3.4.4 Ethyl-substituted pyridines……………… 180 3.4.5 Diazabenzene trends…………………….. 184 3.4.6 Approximations…………………………. 185 3.4.7 Subsequent steps………………………... 186 3.5 Conclusions……………………………………………………...... 188 References………………………………………………………………….. 193

4. Combustion Pathways of the Alkylated Heteroaromatics: Peroxy Radical Pathways of the Five-Membered Heteroaromatics………….. 198

4.1 Introduction…………………………………………………………... 198 4.2 Computational Methods……………………………………………… 203 4.3 Results and Discussion………………………………………………. 205 4.3.1 Furan derivatives……………………………………………. 208 4.3.2 Oxazole derivatives…………………………………………. 218 4.3.3 Pyrrole derivatives…………………………………………... 224 4.3.4 Thiophene derivatives……………………………………….. 227 4.3.5 Overall low-temperature trends……………………………... 230 4.3.6 Heteroatom effects: comparisons to benzylperoxy radical….. 233 4.3.7 Ring size comparison: heteroaromatics……………………... 235 4.3.8 Temperature effects…………………………………………. 236 4.3.9 Ethyl derivatives…………………………………………….. 241 4.3.10 Oscillator vs. rotor approximations………………………… 243 4.4 Conclusions………………………………………………………….. 244 References…………………………………………………………………. 248

ix 5. Combustion Pathways of the Alkylated Heteroaromatics: Reactions of H, O (3P), and HO• with the Alkylated Azabenzenes……………… 252

5.1 Introduction…………………………………………………………... 252 5.2 Computational Methods……………………………………………… 260 5.3 Results and Discussion………………………………………………. 262 5.3.1 Reactions of the picolines……………………………………. 262 5.3.2 Reactions of the alkylated diazabenzenes……………………. 270 5.3.3 Overall trends………………………………………………… 275 5.4 Conclusions…………………………………………………………... 279 References………………………………………………………………….. 281

6. The Multiple Conformations of n-Propylperoxy Radical and Their Implications for Its Unimolecular Decomposition…………………… 286

6.1 Introduction…………………………………………………………... 286 6.2 Computational Methods…………………………………………….... 292 6.3 Results and Discussion………………………………………………. 294 6.3.1 Rotational profiles……………………………………………. 295 6.3.2 Conformer populations………………………………………. 301 6.3.3 Rotational barriers and reaction energetics…………………... 303 6.3.4 Computational approaches…………………………………… 305 6.4 Conclusions…………………………………………………………... 307 References………………………………………………………………….. 309

7. The Unimolecular Decomposition of Propan-1-ol-1-Peroxy Radical: Effects of Functionalization on Peroxy Radical Pathways…………… 313

7.1 Introduction…………………………………………………………... 313 7.2 Computational Methods……………………………………………… 320 7.3 Results and Discussion………………………………………………. 322 7.3.1 Conformations……………………………………………….. 323 7.3.2 Initial H-atom transfers………………………………………. 327 7.3.3 Subsequent decompositions…………………………………. 338 7.3.4 Product distribution and atmospheric implications………….. 342 7.4 Conclusions…………………………………………………………... 346 References………………………………………………………………….. 349

x 8. Non-Aqueous Solvation of n-Octanol and Ethanol: Spectroscopic and Computational Studies………………………………………………... 356

8.1 Introduction…………………………………………………………... 356 8.2 Experimental and Computational Methods………………………….. 360 8.2.1 Raman spectroscopy…………………………………………. 360 8.2.2 Chemicals…………………………………………………….. 360 8.2.3 Computational methods……………………………………… 361 8.3 Results and Discussion………………………………………………. 362 8.3.1 Peak assignments…………………………………………….. 362 8.3.2 Nature of O-H bonds…………………………………………. 366 8.3.3 Computational spectra of ethanol/solvent complexes……….. 367 8.3.4 Rationalization of experimental findings…………………….. 373 8.3.5 Survey of computational methods and basis sets…………….. 374 8.4 Conclusions…………………………………………………………... 377 References………………………………………………………………….. 378

9. The Effects of Hydrogen Augmentation on the Combustion Behavior of Hydrocarbon Fuels: Fuel Blends and Master Equation Analyses…... 383

9.1 Introduction…………………………………………………………... 383 9.1.1 Hydrogen as fuel…………………………………………… 383 9.1.2 Hydrogen as fuel additive………………………………….. 387 9.1.2.1 Generating hydrogen……………………………... 387 9.1.2.2 Methane/H2………………………………………. 388 9.1.2.3 Gasoline/H2………………………………………. 389 9.1.2.4 Diesel/H2…………………………………………. 390 9.1.3 Master equation methods and fuel behavior simulations…... 391 9.2 Computational Methods……………………………………………... 393 9.3 Results and Discussion……………………………………………… 394 9.3.1 Primary reference fuels and hydrogen……………………... 394 9.3.2 n-Heptane/H2 calculations…………………………………. 402 9.4 Conclusions………………………………………………………….. 411 References…………………………………………………………………. 415

Bibliography……………………………………………………………….. 418

xi

LIST OF TABLES

Table Title Page 1.1 Thermodynamic and spin-density information for hydrogen-atom loss reactions of methyl-substituted heteroaromatic rings………………… 63 1.2 Thermodynamic and spin-density information for hydrogen-atom loss reactions of ethyl-substituted heteroaromatic rings…………………... 64 1.3 Thermodynamic information (kcal/mol, 298 K, B3LYP/6- 311+G**//B3LYP/6-31G*) for alkyl group loss reactions of methyl- and ethyl-substituted heteroaromatic rings…………………………… 65 1.4 Comparison of reaction pathway energetics (298 K, kcal/mol) for benzylperoxy radical (B3LYP/6-311+G**//B3LYP/6-31G*)………... 71 2.1 Thermodynamic information and spin density information for hydrogen- loss (C-H homolytic bond cleavage) reactions of methyl-substituted heteroaromatic rings………………………………………. 122 2.2 Comparison of changes in bond lengths (Angstroms) and bond angles (degrees) during the hydrogen-atom loss reactions of toluene and its nitrogen analogue, 2-methylpyridine…………………………………. 123 2.3 Thermodynamic and spin density information for reactions of ethyl- substituted heteroaromatic rings, for hydrogen loss from the CH2 group…………………………………………………………………... 126 2.4 Thermodynamic information for alkyl-group loss reactions of methyl- and ethyl-substituted heteroaromatics………………………………… 128 2.5 Changes in bond lengths (Angstroms) and bond angles (degrees) for hydrogen-loss reaction of representative heteroaromatic ring, 2- methylfuran……………………………………………………………. 133 2.6 Comparison of approximations (harmonic oscillator to hindered rotor), as applied to the thermodynamics of hydrogen-atom loss for methyl- substituted heteroaromatic rings……………………………… 142 3.1 Comparison of DFT (B3LYP/6-311+G**//B3LYP/6-31G*) and CBS- QB3 results for key processes in the reactions possible for benzylperoxy radical………………………………………………….. 161 3.2 Comparison of reaction pathways available to 2-picolinylperoxy radical at 298 K, via B3LYP/6-311+G**//B3LYP/6-31G*…………... 164 3.3 Compiled enthalpies and energies of reaction (kcal/mol) for 3- and 4- picolinylperoxy radical (B3LYP/6-311+G**//B3LYP/6-31G*)……… 170 3.4 Variation in enthalpies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*) for 3- and 4-substituted picolinylperoxy radicals, with increasing temperature……………………………………………….... 179

xii

3.5 Variation in free energies of reaction (B3LYP/6-311+G**//B3LYP/6- 31G*) for 3- and 4-substituted picolinylperoxy radicals, with increasing temperature………………………………………………… 180 3.6 Comparison of free energy profiles (kcal/mol) at 298 K for the picolinylperoxy (methylpyridinylperoxy) and ethylpyridinylperoxy radicals………………………………………………………………… 181 3.7 Free energies of reaction at 298 K (kcal/mol) for reaction pathways available to the peroxy radicals of the diazabenzenes………………… 185 3.8 Free energies of reaction at 298 K (kcal/mol), using rotor approximation in place of harmonic oscillator approximation in calculating the thermal correction to the enthalpy and entropy………. 186 4.1 Free energies of reaction (kcal/mol) with increasing temperature, as calculated via B3LYP/6-311+G**//B3LYP/6-31G*…………………. 211 4.2 Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*) over the 298 to 2000 K range, for peroxy radical rearrangement pathways of the methylated furan and oxazole derivatives……………. 214 4.3 Exploration of multiple heteroatom effect through comparison of the methyloxazolylperoxy derivatives to the corresponding methylfuranylperoxy derivative……………………………………….. 222 4.4 Comparing free energies of activation and reaction for 2-substituted methyl peroxy radical derivatives of furan, pyrrole, and thiophene, relative to starting peroxy radicals……………………………………. 234 4.5 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reaction pathways available to the peroxy radicals of the azabenzenes……………………………………………………………. 236 4.6 Variation in free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298 K) with increasing alkyl substitution (kcal/mol) at 298 K, for the methyl and ethyl-substituted 2- and 3-alkylfuranylperoxy radicals…………………………………………………………………. 242 4.7 Variations in free energies of reaction (B3LYP/6-311+G**//B3LYP/6- 31G*, kcal/mol, 298 K), with varying substitution for the alkylfuranylperoxy, alkylthiophenylperoxy, and alkyloxazolylperoxy radicals. …………………………………………. 243 4.8 Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298 K), using the rigid rotor approximation in place of oscillator approximation……………………………………………….. 234 5.1 Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for hydrogen-atom (H•) reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent ………………………………………………………………… 262 5.2 Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for hydrogen-atom (H•) reactions with the picolines (methyl-substituted pyridines), at each relevant site……………………… 263

xiii 5.3 Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for oxygen atom [O (3P)] reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule……………………………………………………………….. 263 5.4 Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for oxygen atom [O (3P)] reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule………………………………………………………………... 264 5.5 Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for (HO•) reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule……………………………………………………………...... 264 5.6 Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for hydroxyl radical (HO•) reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule…………………………………………………………………. 265 5.7 Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for the H-atom abstraction pathways and addition pathways available to benzene and ……………………………. 269 5.8 Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for the H-atom abstraction pathways and addition pathways available to benzene and pyridine……………………………. 269 5.9 Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reactions of oxygen atom [O (3P)] and the methyl- substituted azabenzenes, at each relevant site within the parent molecule…………………………………………………………………... 270 5.10 Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reactions of oxygen atom [O (3P)] and the methyl- substituted azabenzenes, at each relevant site within the parent molecule…………………………………………………………………. 271 5.11 Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reactions of hydroxyl radical (HO•) and the methyl- substituted azabenzenes, at each relevant site within the parent molecule…………………………………………………………………... 272 5.12 Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reactions of hydroxyl radical (HO•) and the methyl- substituted azabenzenes, at each relevant site within the parent molecule………………………………………………………………….. 273 6.1 Boltzmann distributions for the five rotamers of n-propylperoxy radical, via their free energy differences (kcal/mol, 298 K) and relative degeneracies, as calculated at the CBS-QB3 and hybrid DFT levels of theory……………………………………………………………………… 302

xiv 6.2 Enthalpies and energies of activation and reaction (kcal/mol) for the scission of ethylperoxy radical into ethene and hydroperoxyl radical, calculated via several methods and basis sets…………………………….. 306

6.3 Enthalpies and energies of activation and reaction (kcal/mol) for the cyclization reaction of n-propylperoxy radical, dissociating into cyclopropane and hydroperoxyl radical, calculated via several methods and basis sets……………………………………………………………… 307 7.1 Bond dissociation enthalpies (298 K, kcal/mol, CBS-QB3, B3LYP/6- 31+G**, and mPW1K/ 6-31+G**) for each unique C─H bond in 1- propanol…………………………………………………………………… 322 7.2 Possible dihedral combinations considered in generating starting conformations of propan-1-ol-1-peroxy radical…………………………... 325 7.3 Populations of the five most common rotamers for propan-1-ol-1-peroxy radical, using H0K data calculated via CBS-QB3 and B3LYP/6-31+G** energies and Boltzmann weighting factors……………………………….. 326 7.4 Enthalpies and free energies of reaction (kcal/mol) at 0 K and 298 K for the unimolecular decomposition of propan-1-ol-1-peroxy radical, calculated at the CBS-QB3 level of theory……………………………….. 329 7.5 Enthalpies and free energies of reaction (kcal/mol) calculated at the mPW1K/6-31+G** level of theory at 0 K and 298 K for the unimolecular decomposition of propan-1-ol-1-peroxy radical………………………….. 330 7.6 Enthalpies and free energies of reaction (kcal/mol) calculated at the B3LYP/6-31+G** level of theory at 0 K and 298 K for the unimolecular decomposition of propan-1-ol-1-peroxy radical………………………….. 332 7.7 Comparison of unimolecular decompositions of respective peroxy radicals. CBS-QB3 enthalpies of reaction at 0K and free energies of reaction at 298 K, relative to the global minimum of the respective starting peroxy radical…………………………………………………….. 337 8.1 Boltzmann weighting factors, as calculated from various energy expressions, using B3LYP/6-31+G**//MP2/6-31G* single-point energies……………………………………………………………………. 371 8.2 O−H bond lengths (Å) and stretching frequencies (cm-1) for each distinct complex, as calculated and scaled (factor of 0.9427) at the MP2/6-31G* level of theory……………………………………………………………... 374 8.3 Relative energies (in kcal/mol) compiled using two distinct basis sets, with represented energies relative to that of the lowest-energy complex within each series…………………………………………………………. 376 9.1 Characteristics of hydrogen as compared to typical hydrocarbon fuels…... 380 9.2 Arrhenius coefficients for the combustion of hydrogen. Units in moles, 3 cm , seconds, and K (A, b); kcal/mol (Ea)………………………………... 381

xv

LIST OF FIGURES

Figure Title Page

1.1 High-temperature methane combustion…………………………….. 6 1.2 Low-temperature methane combustion……………………………... 7 1.3 Typical ignition delay of an fuel as a function of the initial mixture’s temperature………………………………………………. 10 1.4 Flow chart for a typical master equation method…………………… 21 1.5 Conformers of n-propylperoxy radical……………………………… 30 1.6 Initial barriers (kcal/mol) for unimolecular reactions of n- propylperoxy radical………………………………………………... 31 1.7 Reactions of toluene with hydroxyl radical…………………………. 37 1.8 High-temperature oxidation pathways of benzene………………….. 39 1.9 (a) Enthalpies (kcal/mol) leading to 2-oxepinoxy radical via unimolecular rearrangement of phenylperoxy radical. (b) Potential decomposition pathways for phenylperoxy radical………..………... 41 1.10 Unimolecular reaction pathways for phenylperoxy radical...... 43 1.11 Unimolecular decomposition pathways of 2-oxepinoxy radical……. 45 1.12 Most favorable oxidative decomposition pathways for 2-oxepinoxy radical at 298 K and 1250 K………………………………………… 46 1.13 Heteroaromatic compounds of interest in modeling coal combustion………………………………………………………….. 52 1.14 Pyrolysis pathway of pyrrole, via intermediate pyrrolenine...... 56 1.15 Unimolecular pathways available to heteroaromatic peroxy radical; example shown for 2-pyridinylperoxy radical……………………… 58 1.16 Cyanocyclopentadiene……………………………………………… 61 1.17 Variation of reaction free energy with temperature for alkyl C-H hydrogen atom loss in (a) five-membered methyl-substituted heteroaromatic rings and (b) six-membered methyl-substituted heteroaromatic rings………………………………………………… 67 1.18 Unimolecular pathways for 2-picolinylperoxy radical, a representative peroxy radical for alkylated azabenzenes……………. 69 1.19 ΔG298 of reaction versus temperature for the unimolecular pathways of 2-picolinylperoxy radical…………………………………………. 73 1.20 ΔG298 of activation versus temperature for the unimolecular pathways of 2-picolinylperoxy radical………………………………. 74

xvi 1.21 Qualitative depiction of favorable cyclization pathways for representative peroxy radicals of methyl heteroaromatics…………... 76 1.22 Initial reaction steps for fatty decomposition………………….. 83 2.1 Heteroaromatic rings of interest……………………………………... 106 2.2 Generic representation of possible radical-generation pathways for the alkylated heteroaromatic rings…………………………………... 111 2.3 Comparison of areas of increased spin densities (α-β) for 2- methylpyridinyl radical and 2-ethylpyridinyl radical………………... 120 2.4 Variation of bond dissociation enthalpy (kcal/mol) with excess spin density at the incipient CH2 radical center, for methyl-substituted heteroaromatic rings…………………………………………………. 126 2.5 Variation of bond dissociation enthalpy (kcal/mol) with excess spin density at the incipient C-H radical center, for ethyl-substituted heteroaromatics……………………………………………………… 127 2.6 Correlation between BDEs for methyl and ethyl-substituted derivatives of the five-membered heteroaromatics...... 128 2.7 Variation of bond dissociation enthalpy (kcal/mol) with temperature (K) for hydrogen-atom loss in the (a) five-membered methyl- substituted heteroaromatics and (b) six-membered methyl- substituted heteroaromatics………………………………………….. 131 2.8 Variation of bond dissociation enthalpy (kcal/mol) with temperature (K) for hydrogen-atom loss in the (a) five-membered ethyl- substituted heteroaromatics and (b) six-membered ethyl-substituted heteroaromatics……………………………………………………… 132 2.9 Variation of free energy of reaction (kcal/mol) with temperature (K) for hydrogen-atom loss in the (a) five-membered methyl-substituted heteroaromatic rings and (b) six-membered methyl-substituted heteroaromatic rings…………………………………………………. 133 2.10 Variation of free energy of reaction (kcal/mol) with temperature (K) for hydrogen-atom loss in the (a) five-membered ethyl-substituted heteroaromatic rings and (b) six-membered ethyl-substituted heteroaromatic rings………………………………………………… 134 2.11 Comparison of hydrogen-atom loss and methyl-group loss free energies with increasing temperature………………………………... 135 2.12 Comparison of hydrogen-atom loss and ethyl-group loss free energies with increasing temperature………………………………... 136 3.1 Heteroaromatic rings of interest (the azabenzenes)…………………. 146 3.2 Likely pathways for the alkylated aromatic peroxy radicals, using benzylperoxy radical as a model compound………………………… 150 3.3 Numbering scheme for the azabenzylperoxy radicals……………….. 158 3.4 Change in bond lengths with addition of nitrogen atom to the aromatic system……………………………………………………… 161 3.5 Variation in excess spin density via cyclization at the ortho position, for benzylperoxy radical and 2-picolinylperoxy radical…………….. 163

xvii 3.6 H-transfer pathways delineated for each of the three picolinylperoxy radicals………………………………………………………………. 166 3.7 Variation in ΔHrxn with temperature for the pathways available to 2- picolinylperoxy radical………………………………………………. 169 3.8 Variation in ΔGrxn with temperature for the pathways available to 2- picolinylperoxy radical………………………………………………. 170 3.9 Change in ΔHactivation with increasing temperature, for 2- picolinylperoxy radical………………………………………………. 172 3.10 Change in ΔGactivation with increasing temperature, for 2- picolinylperoxy radical………………………………………………. 173 3.11 Subsequent steps available to certain peroxy radical rearrangement products…………………………………………………………….... 183 4.1 Common five-membered heteroaromatic rings and their numbering schemes………………...... 199 4.2 Rearrangement pathways available to a typical peroxy radical derivative of an alkylated five-membered heteroaromatic ring……... 206 4.3 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 2-methylfuranylperoxy radical, shown relative to the starting peroxy radical………………... 209 4.4 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 3-methylfuranylperoxy radical (4.2), shown relative to the starting peroxy radical………….. 210 4.5 Allylic stabilization afforded via the most likely cyclization pathways of 2-methylfuranylperoxy radical (top) and 3- methylfuranylperoxy radical (bottom)……………………………… 216 4.6 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 2-methyloxazolylperoxy radical, shown relative to the starting peroxy radical………………... 219 4.7 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 4-methyloxazolylperoxy radical, shown relative to the starting peroxy radical………………... 220 4.8 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 5-methyloxazolylperoxy radical, shown relative to the starting peroxy radical………………... 221 4.9 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 2-methylpyrrolylperoxy radical, shown relative to the starting peroxy radical………………... 224 4.10 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 3-methylpyrrolylperoxy radical, shown relative to the starting peroxy radical………………... 225 4.11 Free energies of potential steps (B3LYP/6-311+G**//B3LYP/6- 31G*, kcal/mol, 298 K) available to the N-centered radical formed via ortho cyclization / hydrogen-atom transfer in 2- methylpyrrolylperoxy radical………………………………………... 227

xviii 4.12 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 2-methylthiophenylperoxy radical, shown relative to the starting peroxy radical………………... 228 4.13 Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 3-methylthiophenylperoxy radical, shown relative to the starting peroxy radical………………... 229 4.14 Comparison of 2-furanylperoxy radical (A) to 2- methylfuranylperoxy radical (B)…..………………………………… 232 4.15 Peroxy radical pathways resulting in aryl or aryloxy radicals either directly or indirectly, along with pathway resulting in carbonyl- functionalized compound……………………………………………. 237 4.16 Variation in ΔGrxn (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol) with temperature for the pathways available to 2- methylfuranylperoxy radical………………………………………… 239 4.17 Variation in ΔGactivation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol) with temperature for the most energetically favorable pathways available to 2-methylfuranylperoxy radical………………. 240 5.1 Possible pathways for the reaction of hydroxyl radical (HO•) with 2- methylpyridine………………………………………………………. 254 5.2 (a) The azabenzenes. (b) The picolines (methylpyridines) and toluene, model compounds for understanding the chemistry of the azabenzenes with H•, O (3P), and HO•………………………………. 258 5.3 Representation of ring-opening reactions unique to methylpyridazines, when atomic oxygen attacks ortho to the nitrogen atoms……………………………………………………….. 274 5.4 Correlation between the bond dissociation enthalpies (kcal/mol) for hydrogen-atom loss at the of the alkylated heteroaromatic rings and the corresponding free energy of activation (kcal/mol) for H-atom abstraction at the methyl group by HO•……... 276 5.5 Correlation between the bond dissociation enthalpies (kcal/mol) for hydrogen-atom loss at the methyl group of the alkylated heteroaromatic rings and the corresponding free energy of activation (kcal/mol) for H-atom abstraction at the methyl group by O (3P)…... 277 5.6 Relationship between the free energy of activation (kcal/mol) for H- atom abstraction by HO• at a given position and the free energy of activation (kcal/mol) for addition of HO• to that same position…….. 278 6.1 n-Propylperoxy radical, wherein 1,5-H atom transfer can readily occur via a six-membered ring transition state………………………. 281 6.2 Conformeric possibilities for n-propylperoxy radical……………….. 285 6.3 Rotational profile for C2−C3 bond of n-propylperoxy radical (B3LYP/6-31+G**), as a function of electronic energy (kcal/mol)...... 289 6.4 Rotational profile for C2−C3 bond of n-propylperoxy radical (mPW1K/6-31+G**), as a function of electronic energy (kcal/mol)... 290

xix 6.5 Visual representations of highest-energy rotational conformer for rotation around the C−C−C−O dihedral in n-propylperoxy radical…. 291 6.6 Rotational profile for the C3−O bond of n-propylperoxy radical (B3LYP/6-31+G**), as a function of electronic energy (kcal/mol)… 292 6.7 Rotational profile for the C3−O bond of n-propylperoxy radical (mPW1K/6-31+G**), as a function of electronic energy (kcal/mol)... 293 6.8 Visual representations of highest-energy rotational conformer for rotation around the C−C−O−O dihedral in n-propylperoxy radical…. 293 6.9 Rotational profile for methyl rotation in n-propylperoxy radical, (B3LYP/6-31+G**, kcal/mol)………………………………………. 294 6.10 Initial barrier heights for unimolecular decomposition of n- propylperoxy radical, expressed in terms of ΔH(298 K), kcal/mol, relative to n-propylperoxy radical…………………………………… 298 7.1 Common steps of peroxy radical chemistry…………………………. 309 7.2 Propan-1-ol-1-peroxy radical………………………………………... 317 7.3 Likeliest conformers of propan-1-ol-1-peroxy radical………………. 321 7.4 Isomerization pathways available to propan-1-ol-1-peroxy radical, based on previous work with n-propylperoxy radical……………….. 322 7.5 Initial kinetic barriers for pathways available to propan-1-ol-1- peroxy radical, expressed in terms of ΔH0K (kcal/mol), calculated via CBS-QB3 and DFT (B3LYP/6-31+G** and mPW1K/6- 31+G**) methods……………………………………………………. 328 7.6 Overall potential energy surface for the decomposition of propan-1- ol-1-peroxy radical using ΔH0K (kcal/mol) values calculated at the CBS-QB3 level of theory……………………………………………. 333 8.1 Raman spectra of neat octanol and ethanol………………………….. 362 8.2 Experimental Raman spectra of (a) neat n-octanol, and 0.05x n- octanol in (b) benzene, (c) cyclohexane, and (d) carbon tetrachloride, in the 3550 to 3700 cm-1 region of the free O−H stretch………………………………………………………………... 364 8.3 Experimental Raman spectra of (a) neat ethanol, and 0.05x ethanol in (b) benzene, (c) cyclohexane, and (d) carbon tetrachloride, in the 3550-3700 cm-1 region of the free O−H stretch……………………... 365 8.4 Ethanol:benzene complexes as calculated at the HF/6-31G* (top) and MP2/6-31G* (bottom) levels of theory…………………………. 368 8.5 Ethanol:cyclohexane complexes as calculated at the HF/6-31G* (top) and MP2/6-31G* (bottom) levels of theory…………………… 368 8.6 Ethanol:CCl4 complexes as calculated at the HF/6-31G* (top) and MP2/6-31G* (bottom) levels of theory……………………………… 369 8.7 Gas-phase (a) and PCM (b) calculated Raman spectra, generated via MP2/6-31G* calculations and Boltzmann weighting factors from H0 values (from B3LYP/6-31+G**//MP2/6-31G* energies)…………… 370 9.1 Generalized depiction of the progress of a master equation method... 392

xx 9.2 Variation in flame speed with hydrogen augmentation to PRF combustion mechanism, as predicted by CHEMKIN 4.1…………… 395 9.3 Variation in hydrocarbon fuel consumption (top, iso-octane; bottom, n-heptane) with H2 augmentation, as predicted by PRF mechanism... 396 9.4 Variation in H2 mole fraction as predicted by PRF mechanism, as a function of H2 augmentation (top) and of temperature change (bottom)……………………………………………………………… 397 9.5 Variation in emissions with H2 augmentation to PRF mechanism……………………………………………………… 398 9.6 Variation in unburned fuel fraction with H2 augmentation to PRF mechanism…………………………………………………………… 398 9.7 Variation in heat production with H2 augmentation to PRF mechanism…………………………………………………………… 399 9.8 Non-linear variation in flame speed at a fixed point (0.15 cm), for H2 augmentation of PRF mechanism………………………………... 400 9.9 Relative overall trends with H2 augmentation to PRF mechanism….. 401 9.10 Variation in flame speed with H2 augmentation to n-heptane mechanism (Φ = 1.2)………………………………………………… 403 9.11 Variation in flame speed at a fixed point (2.4 cm) for H2 augmentation of n-heptane (Φ = 1.2)………………………………... 404 9.12 Variation in n-heptane consumption with increasing H2 augmentation………………………………………………………… 404 9.13 Variation in H2 concentration with increasing H2 augmentation……. 405 9.14 Variation in carbon monoxide emissions with increasing H2 augmentation………………………………………………………… 405 9.15 Variation in net heat production with H2 augmentation to n-heptane.. 406 9.16 Variation in unburned fuel fraction with H2 augmentation to n- heptane (Φ = 1.2)....…………………………………………………. 406 9.17 Relative trends associated with H2 augmentation for n-heptane mechanism (Φ = 1.2)………………………………………………… 407 9.18 Variation of flame speed with equivalence ratio for n-heptane, as calculated experimentally and theoretically…………………………. 409 9.19 Linear variation in fuel-rich (Φ = 2.0) flame speed with H2 augmentation to n-heptane…………………………………………... 410 9.20 Relative trends with H2 augmentation for fuel-rich n-heptane mechanism (Φ = 2.0)………………………………………………… 411

xxi

CHAPTER 1

INTRODUCTION: THE CHEMISTRY OF REACTIVE RADICAL INTERMEDIATES IN COMBUSTION AND THE ATMOSPHERE

1.1. Introduction

Combustion processes convert chemical energy into heat energy, playing several

important roles in today’s society. Oxidation processes provide power to beneficiaries

ranging from automobiles to electrical generators; atmospheric oxidation reactions

impact a wide range of environmental phenomena (i.e., formation, photochemical

smog, and acid rain). To fully understand combustion chemistry, it is necessary to

understand the properties of the common reactive intermediates that participate in these

reactions. Alkyl radicals (R•), alkoxy radicals (RO•), and peroxy radicals (ROO•) constitute the main classes of reactive radical intermediates involved in combustion. The occurrence and stability of the intermediates is governed by the temperature and pressure at which reactions take place. Understanding these complex and dynamic relationships presents challenges for experimentalists and also for theorists. This chapter will first examine the fundamental chemistry that occurs during combustion of a fuel, and then move into an exploration of some key reactive intermediates. In particular, we will focus

1 on the importance of to combustion processes, highlighting our

work on the unimolecular dissociative pathways available to the peroxy radicals of alkyl,

aromatic, and heteroaromatic compounds.

1.2. Basic Concepts of Combustion Chemistry

1.2.1. Free radicals. Any type of combustion chemistry is essentially dictated by the

radical intermediates present. Before discussing how this occurs, it is first instructive to

review some key aspects of radical chemistry. Most simply, free radicals are

with unpaired electrons. The lifetimes of these species vary widely given molecular

composition and reaction environment, but the alkyl radicals of interest in combustion

chemistry are highly reactive and thus short-lived. Radicals can vary in their

hybridization (sp3, sp2, or sp), as well as the nature of the orbital in which their unpaired

electron is placed – one dominated by s or p character – and thereby dictating the

geometry around the radical center. Furthermore, a free radical is stabilized by alkyl

substituents on the radical center; tertiary radicals are more stable than secondary

radicals, which are correspondingly more stable than primary radicals. Resonance

stabilization (electron delocalization) also plays a role in radical stability: a radical

adjacent to a π network (i.e., an allylic or benzylic system) can delocalize the unpaired

electron through this system and gain stability.

A radical is an odd-electron molecule, due to an unpaired electron, which can be oriented either spin-up () or spin-down (), leading to two degenerate electronic states as a doublet. This electronic configuration enables reactions that are different from those

2 of closed-shell molecules. The basic steps of a radical chain reaction are familiar from

undergraduate organic chemistry: initiation, whereby a reactive radical species forms;

propagation, the processes by which the radicals react with other molecular species to

generate other radicals; and termination, via the collision of any two radical species to

form one closed-shell molecule, thereby removing radicals from the system. Initiation

can occur either thermolytically (when heat homolytically breaks a molecule’s bond) or

photolytically (when high-energy light homolytically breaks a molecule’s bond). Once a

radical forms, it can undergo a variety of propagation reactions, including hydrogen-atom

transfers, eliminations, additions, unimolecular fragmentations; any reaction step that

begins with a radical reactant and yields a radical product is classified as a chain-

propagating step. Often, in flame chemistry, a reactive hydroxyl (HO•) radical is first

formed, which then reacts with the fuel molecule via an initiation step (R–H + HO• → R•

+ H2O). The ensuing, various propagation possibilities constitute much of the chemistry

of interest in combustion processes; this topic will be revisited shortly.

Bond dissociation enthalpies (BDEs) aid in predicting relative reactivities for

different organic (fuel) molecules. BDEs correspond to the enthalpy change for the

homolytic cleavage of a chemical bond. The more stable the resulting radical is, relative

to the reactant, the more likely the bond is to break. General explanations for radical

stability include alkyl substitution and resonance effects (as noted above); however,

Gronert has recently suggested an alternative explanation, that the potential for release of

1,3 repulsive energy (strain) has the greatest effect on these quantities: i.e., a tertiary

radical is more stable than a primary radical not because of the larger number of alkyl

3 substituents attached to the radical center, but because of the greater magnitude of the

geminal interactions in the parent (more congested) hydrocarbon that are relieved upon

C−H bond cleavage to form the radical.1 Regardless of the origin of the effect, these

BDE quantities are useful in rationalizing initiation steps for given fuels.

Free radicals are involved in a wide variety of reactions, due to their reactivity

and versatility; understanding their behavior as it relates to combustion is a goal of many

experimentalists and theoreticians. Two specific classes of radicals of interest to combustion processes are peroxy (ROO•) radicals and oxy (RO•) radicals; along with

hydroxyl (HO•) radical, many of these radicals are important members in the general

class referred to as reactive oxygen species (ROS).2

1.2.2. Combustion at varying temperatures. The concept behind combustion is

straightforward—when a hydrocarbon fuel reacts with oxygen, the organic component is eventually converted to and —but the reality is more complicated.

For instance, the combustion of methane (Reaction 1.1) is often used to teach students how to balance reaction equations:

CH4 + 2 O2 → CO2 + 2 H2O (1.1)

The combustion process for methane requires no fewer than 325 individual mechanistic

steps to be accurately described, rather than the one-step route shown above.3 As such,

incomplete combustion is a common occurrence and reactive oxygen species are

pervasive byproducts of that phenomenon, affecting fuel efficiency, and producing

atmospherically detrimental emissions. Moreover, combustion varies with system

4 temperature, as different oxidative pathways become accessible. By examining the representative cases of methane oxidation over a range of temperatures, this phenomenon becomes clearer.

1.2.2.1. High-temperature combustion. At high temperatures (T > 1000 K), methane oxidation is initiated via hydrogen-atom abstraction by hydroxyl radical, oxygen atom, or hydrogen atom (all of which are species generated in flames). Subsequently, in the most direct oxidative route (Figure 1.1), is oxidized to , which then loses a hydrogen atom to form formyl (HCO•) radical.

5

Figure 1.1: High-temperature methane combustion.

Formyl radical can lose a subsequent hydrogen atom via collisional dissociation or reaction with molecular oxygen, thereby forming carbon monoxide (CO). A final oxidation step via the reaction of CO with hydroxyl radical yields the fully oxidized carbon dioxide as a final product. To fully characterize the overall methane combustion process, additional reactions such as methyl radical recombination and hydrocarbon eliminations must also be considered. In this basic scenario, methyl, methoxy, and formyl radicals are the intermediates of greatest interest; extending beyond this most straightforward case, the reactions of alkyl, alkoxy, and aldehydic radicals are all reactive intermediates of interest in high-temperature fuel combustion.

1.2.2.2. Low-temperature combustion. The low-temperature oxidation of methane (T <

1000 K) presents a more complex reaction scheme (Figure 1.2).

6

Figure 1.2: Low-temperature methane oxidation.

It can be rationalized as proceeding in two phases.4 Since reactive flame species such as

O(3P), H•, and HO• are not observed at low temperatures, the formation of the initial methyl radical must be achieved instead via an endothermic reaction with molecular oxygen.5 Once methyl radical is formed, it reacts with another oxygen molecule to form

• methylperoxy radical (CH3OO ), which abstracts a hydrogen atom from methane to form methyl (CH3OOH) and methyl radical. Methyl hydroperoxide can

• • unimolecularly dissociate to produce methoxy (CH3O ) and hydroxyl (HO ) radical.

These last two steps are crucial, as they build up a reactive radical pool. Once a

sufficient amount of methyl, methoxy, and hydroxyl radicals has formed, these species

7 appropriate the initial duty of abstracting H atoms from methane, driving the reaction rate

forward rapidly. Methyl radical is now formed at an appreciable rate and can undergo

oxidation and recombination steps.

More generally, low-temperature combustion relies heavily on the tendency of

radical propagation to yield chain-branching reactions, a phenomenon first explored by

Semenov.6 Semenov’s reaction scheme is most relevant for species with two or more and can be written as:

• • R + O2 → + HO2 (1.2)

• • R + O2 + M → RO2 + M (1.3)

• • RO2 + R−H → RO2H + R (1.4)

• • RO2 → R'CH(=O) + R"O (1.5)

• • HO2 + R−H → H2O2 + R (1.6)

• • RO2H → RO + OH (1.7)

• • R'CH(=O) + O2 → R'C (=O) + HO2 (1.8)

Using this scheme, we can track the original radical through the most likely mechanisms for oxidation at low temperatures that lead to chain-branching. Once formed from the parent molecule (R−H), an alkyl radical (R•) can react with molecular oxygen to form an

• alkene and hydroperoxyl (HO2 ) radical (Equation 1.2), via 1,4-H-atom abstraction.

Alternatively, methylperoxy radicals can be formed following collisional stabilization by

another molecule or atom (M). The branching ratio for these two reactions is highly

pressure-dependent (and thus depends on concentration of M). Once formed, the peroxy

radical can abstract a hydrogen atom from an alkane to form a new radical (Equation 1.4)

8 or dissociate unimolecularly, yielding an and alkoxy radical (Equation 1.5); both of these reactions produce a molecule that can participate in a chain-branching

• reaction. Since HO2 is unreactive at lower temperatures, reaction 1.6 is less likely than

• the self-reaction of HO2 resulting in H2O2 and O2. Reactions 1.7 and 1.8 are chain- branching steps, which figure heavily in the exponential increase in radical concentration necessary to achieve ignition for the combustion of a given fuel (R−H).

As shown, peroxy radical chemistry plays a substantial role in low-temperature processes, as opposed to the alkoxy radical chemistry of high-temperature combustion.

Thus, the peroxy radicals constitute an important class of reactive intermediates with significant implications for combustion and atmospheric reactions.

1.2.2.3. Negative temperature coefficient (NTC) phenomenon. In terms of temperature regions, low-temperature combustion occurs over the range 298–550 K, whereas high- temperature combustion mechanisms dominate at temperatures over 1000 K.

Intermediate temperatures, from 550 to 700 K, demonstrate an unusual phenomenon called the negative temperature coefficient (NTC), which is observed for methane and larger hydrocarbon fuels.7 As shown in Figure 1.3, when the correct alkylperoxy radical chemistry is included in a fuel’s combustion mechanism, a NTC range exists (Figure 1.3, plot C) and an increase in temperature causes a decrease in reaction rate (i.e. a longer time to ignition).8 This phenomenon can be rationalized via an examination of the low- temperature combustion mechanism.

9

Figure 1.3: Typical ignition delay of an alkane fuel as a function of the initial mixture’s temperature. Three different kinetic models are shown: (a) High temperature chemistry only; i.e., no peroxy radical chemistry. (b) Same as (a), but the “Q•OOH” chain- branching channel of the peroxy radicals has been considered. (c) Same as (b), but the • • concerted elimination of RO2 to olefin + HO2 has been considered. (Figure courtesy of Dr. Timothy Barckholtz, ExxonMobil Research and Engineering.)

Again, a crucial step in achieving the combustion of hydrocarbon fuels involves

the unimolecular dissociation of an alkyl hydroperoxide (ROOH) into alkoxy (RO•) and hydroxyl (HO•) radicals, i.e., chain-branching. The alkylperoxy radical leading to an alkyl hydroperoxide (Equation 1.4) is in equilibrium with the reactants, alkyl radical and molecular oxygen (O2) (reverse of Equation 1.3). As the temperature increases, entropy

favors the reactants, so that the alkylperoxy radical concentration will be minimized, and

therefore, so will the alkyl hydroperoxide concentration. Thus, the chain-branching step

cannot drive the oxidation forward, and the ignition time will increase. Moreover, for a

hydrocarbon with two or more carbons, molecular oxygen will instead abstract a

10 hydrogen atom from the alkyl radical, yielding an alkene and (notoriously unreactive)

• hydroperoxyl (HO2 ) radical (Equation 1.2). The NTC regime will persist until the temperature increases sufficiently to allow for high-temperature, chain-branching pathways. At high-temperatures, NTC is no longer relevant and the ignition rate increases with increasing temperature once again.

1.2.2.4. Atmospheric oxidation. The chemistry of the troposphere overlaps with low-

temperature combustion, as one would expect for an oxidative environment.

Consequently, the concerns of atmospheric chemistry overlap with those of combustion

chemistry. Monks recently published a tutorial review of radical chemistry in the

troposphere,9 while Atkinson and Arey have compiled a thorough database of

atmospheric degradation reactions of volatile organic compounds, which addresses these

issues in far greater detail than is possible for the scope of this chapter.10 We can extend the low-temperature combustion reactions (Equations 1.4 and 1.5), whereby peroxy radicals would have the capacity to react with prevalent atmospheric radicals, such as

NO, NO2, and NO3 (collectively known as NOx):

• • • • RCH2O2 + NO → RCH2O + NO2 (1.9)

• • RCH2O2 + NO + M → RCH2ONO2 + M (1.9a)

• • RCH2O2 + NO2 + M → RCH2OONO2 + M (1.10)

• • RCH2O2 + HO2 → RCH2OOH + O2 (1.11)

• • RCH2O2 + HO2 → RCHO + H2O + O2 (1.11a)

• • • • RCH2O2 + RCH2O2 → RCH2O + RCH2O + O2 (1.12)

11 • • RCH2O2 + RCH2O2 → RCHO + RCH2OH + O2 (1.12a)

• • • • RCH2O2 + NO3 → RCH2O + NO2 + O2 (1.13)

Oxidation in the atmosphere begins photolytically with radiation from the sun

rather than thermolytically; thus, atmospheric chemistry differs between day and night.

In the daytime, the most common initiation step involves photolysis of ozone by the sun’s ultraviolet light, leading to hydroxyl (HO•) radical generation:

1 1 O3 + hυ → O( D) + O2( Δg) (1.14)

1 • O( D) + H2O → 2 HO (1.15)

Once hydroxyl radical is formed, it can abstract hydrogen atoms from (in

the atmosphere, a hydrocarbon is generally referred to as a volatile , or

VOC) to generate alkyl radicals, which can react similarly to those in low-temperature

combustion. VOCs are emitted into the atmosphere via both natural and anthropogenic

(man-made) processes. Again, since peroxy radical concentrations are pressure-

dependent, these species can revert back to reactants if they are not collisionally

stabilized. Once formed, the peroxy radicals have longer lifetimes than HO• radical and

maintain larger concentrations in the atmosphere; common reactions of peroxy radicals include self-reactions to yield two carbonyl-functionalized species and O2 (Equations

• 1.12 and 1.12a); reactions with HO2 to yield alkylhydroperoxides, , H2O, and

O2 (Equations 1.11 and 1.11a); and reactions with NOx species to yield alkoxy radical and

moieties. NOx reactions are of significant interest for these latter species because

• • they decompose into RO and NO2 . Alkoxy radicals can react with oxygen to yield

aldehydes; moreover, after H-atom transfer, the weak aldehydic C−H bond is readily

12 • abstracted to yield an acyl radical that can react with O2 or NO2 in succession to form

peroxyacylnitrates (RC(=O)ONO2 or PANs). PANs have various detrimental effects; they are lachrymators and demonstrate mutagenic effects.11 Moreover, these PANs are

long-lived and can act as reservoirs for NOx and VOCs, traveling far from their points of origin, thus increasing their lifetime and geographical impact.

At night, when the sun’s radiation is minimal, the dominant oxidant is nitrate

• • • radical (NO3 ). This oxidant’s chemistry differs from HO radical in that NO3 prefers to react with unsaturated compounds via addition to one of the carbons of the π-system, rather than by hydrogen-atom abstraction:

• • NO2 + O3 → NO3 + O2 (1.16)

• • NO3 + CH3CH=CH2 → CH3 CH═CH2ONO2 (1.17)

• • CH3 CH═CH2ONO2 + O2 → CH3CH(OO )CH2ONO2 (1.18)

• • As shown, NO3 radical leads to different chemistry than does HO radical; the peroxy

radical can decompose to yield several products, including , formaldehyde,

1,2-propanediol dinitrate (PDDN), nitroxyperoxypropyl nitrate (NPPN) and α-

nitrooxyacetone. The reactions of the peroxy radicals with NOx species can lead to

highly functionalized (and oxidized) organic compounds.

• The interplay of HO , peroxy radicals, VOCs, and NOx species has substantial

implications for tropospheric air quality. For instance, VOCs, NOx, and sunlight result in

poor visibility from ozone as well as aerosol formation, together denoted as

photochemical smog, which can lead to health effects in sensitive individuals. Normally,

we think of reductions in either class of compounds as beneficial to the atmosphere.

13 However, reducing VOC emissions only impacts ozone concentration in high-NOx areas.

Moreover, in VOC-sensitive areas, reductions in NOx may lead to the overproduction of

ozone. We can examine a simplified scheme12 for ozone production:

• • 3 NO2 + hυ → NO + O( P) (1.19)

3 O( P) + O2 + M → O3 + M (1.20)

• • In an ideal troposphere, O3 would react with NO yielding NO2 , and ultimately

regenerating O3 (Equations 1.19 and 1.20), thus no over-production of ozone could occur.

However, in the presence of VOCs, the resultant peroxy radicals formed can compete

• • with O3 and instead react with NO to form excess NO2 , thus resulting in the formation

of excess ozone. This is just one example of the complexity of atmospheric chemistry;

peroxy radicals and NOx have substantial implications for reaction with climate gases,

acid rain formation, and other aspects of air quality. Saunders et al.13 and Jenkin et al.14 have provided a wealth of information on the tropospheric degradation of aliphatic and aromatic VOCs. Additionally, the interested reader may wish to consult References 7-9 for further discussion of these important topics.

An additional concern in the commercial applications of combustion chemistry involves understanding and minimizing the production of harmful emissions (such as carbon monoxide, carbon dioxide, and the oxides of sulfur (SOx), in addition to NOx) in combustion processes. Carbon dioxide increases the amounts of greenhouse gases in the atmosphere and contributes to global warming, while NOx and SOx (oxidized derivatives

of nitrogen and sulfur) can be transformed in water aerosols resulting in acid rain (via

HNO3 and H2SO4, respectively). Another concern involves the formation of soot

14 particles that can have severe respiratory effects. Many studies in combustion chemistry

seek to reduce the production of these harmful byproducts.15

1.2.3. Methods for studying reactive combustion species

1.2.3.1. Experimental methods. Experiments performed in the 1960s and 1970s used a

variety of approaches to detect molecular species and determine rate coefficients for the elementary steps that comprised hydrocarbon combustion. Many of these experimental methods are still widely used in combustion studies today. We will briefly discuss these methods as well as surveying some of the more sophisticated methods developed in recent decades.

In combustion experiments, there are two key considerations: first, generating a

flame; and second, detecting the species of interest. Gaseous flows in a flame can be classified as laminar (streamlined layers) or turbulent. While these flames can be analyzed directly, it is less confounding to study flame chemistry through controlled

generation of reactive species in one of a wide variety of experimental apparata.

One such device is the shock tube.16 This cylindrical apparatus has both a high- pressure region filled with an inert gas and a low-pressure region that contains the reactants of interest—fuel only if pyrolytic processes are being examined, fuel and oxidizer if combustion processes are being examined— separated by a thin membrane.

By rupturing the membrane, a high-pressure shock wave travels down the low-pressure region and is reflected back on a microsecond time scale. The system temperature

15 increases rapidly enough to yield high-temperature chemistry and is rapidly quenched for

clean analysis.

Other techniques simulate post-ignition flame processes. In the flow reactor,

reactants of interest enter the reactor at one end and travel through a constant-temperature region. Ideally, all concentrations and temperatures are consistent across the cross- section of the reactor, so that all movement is in one direction and wall reactions are minimized. A similar approach involves the use of crossed molecular beams,17 wherein two molecular beams are directed into one another; the area of collisional intersection demonstrates chemistry that can occur in flames.

Additionally, specific flame species of interest can be directly generated.

Radicals can be generated either photolytically (via light) or thermolytically (via heat).

Laser-based methods are used to photolytically generate radicals. However, not every

radical of interest can be generated from a convenient precursor; moreover, radicals

generated via photolytic methods have excess internal energy, which increases the

potential of their side reactions. Likewise, pyrolytic sources can be used with a wider

range of species, but often necessitate a long residence time (ms) for radicals in the

heating zone, giving high probabilities for radical-radical reactions and radical-wall

reactions. These drawbacks have been circumvented via novel instrumentation in certain

cases. For instance, Chen et al. developed a hyperthermal pyrolytic nozzle18 that works

on a shorter time scale (μs) and generates thermally cold radicals via a jet expansion; this

nozzle has been coupled with several targets and techniques by various researchers.19

16 In terms of coupling flame generation and detection methods, several

combinations are common. Generally, shock tubes are coupled with IR and UV

absorption and gas chromatography (GC) detectors, while flow reactors are used in

tandem with GC, electron spin resonance, and resonance fluorescence detection.

More advanced techniques are also available. For instance, mass spectrometry

(MS) can be used by converting neutral radicals to ions, via chemical ionization and

electron impact; these ions are then separated and detected according to their mass-to-

charge ratios (m/z). The primary drawback of this method is that it cannot directly

discern between isomers; structural isomers have identical m/z ratios, but may behave

unique chemically. Moreover, when MS is coupled with a flame-generating system, the

probe location must be chosen carefully, as the species population can vary greatly given

the distance from the flame.

Instrumental methods have become more sophisticated to face these challenges.20

In particular, Westmoreland and Cool have developed a flame-sampling mass spectrometer that has provided several revelations in terms of relevant molecular intermediates in combustion.21 Their set-up couples a laminar flat-flame burner to a mass

spectrometer. This burner can be moved along the access of the molecular beam, to obtain spatial and temporal profiles of common flame intermediates. By using a highly tunable synchrotron radiation source, isomeric information on selected mass peaks can be obtained. This experiment represents a huge step forward in the utility of MS in combustion studies: lack of isomer characterization had previously prevented a full accounting of the reaction species and pathways.

17 When paired with an appropriate radical target, laser-based methods can serve

either diagnostically, to discern which intermediates are present in a flame, or

analytically, to explore the kinetics and dynamics of elementary steps of interest. Laser-

induced fluorescence (LIF) and Raman spectroscopy (RS) are two typical diagnostic

techniques; the former explores electronic transitions of a radical of interest, while the

latter explores structural aspects of a species via observation of changes in its

polarizability. Laser flash photolysis is a common kinetic technique, in which a radical

precursor is generated via a laser pulse, and the resultant radicals react with a target of

interest; the rate of this reaction can be thus extrapolated by monitoring decreasing radical concentration over time. Many laser techniques are currently available and can provide a wealth of information on the structural, vibrational, and electronic properties of reactive radical intermediates. Several reviews of the use of lasers in combustion chemistry have been compiled, including those by Wolfrum,22 Crosley,23 and Eckbreth.24

1.2.3.2. Computational methods. Computational and experimental methods clearly

benefit from a symbiotic relationship in combustion studies.25 Theoretical calculations

can propose potential pathways yielding empirically-observed intermediates by providing

reaction energies and rate coefficients of elementary reactions, thereby guiding

experiments. Moreover, theoretical calculations can supplement some gaps caused by

limitations in experimental approaches: the vast majority of analytical techniques fail to

distinguish between structural isomers and to identify discrete intermediate species, both

18 of which are important objectives in delineating overall combustion behavior. Finally,

modeling can identify species to look for experimentally.

Quantum chemical calculations are the most accurate theoretical methods

available for studying the structures, energies, and elementary reactions for molecules. It

is possible to determine the structure, energy, and geometrical parameters (i.e.,

vibrational frequencies, electronic states, rotational constants) for reactants, transition

states and products of a chemical reaction. With this information, the enthalpy (H),

entropy (S), and free energy (G) at specified temperatures can be estimated and reaction barriers and energies predicted. Furthermore, since we can calculate reaction barriers,

absolute reaction rate coefficients can be calculated using transition state theory (TST).26

Tk ≠ Γ= TTk )()( B e Δ− 0 BT)/kG( (1.21) TST h

≠ T is the absolute temperature, h is Planck’s constant, kB B is Boltzmann constant, and ΔG 0 is the free energy barrier height relative to reactants at infinite separation. The temperature dependent factor Γ(T) represents quantum mechanical tunneling and the

Wigner approximation27 to tunneling through an inverted parabolic barrier:

2 1 ⎛ hvi ⎞ T 1)( +=Γ ⎜ ⎟ (1.22) 24 ⎝ BTk ⎠

where νi is the imaginary vibrational frequency representing the TS barrier’s curvature.

Transition state theory yields rate coefficients at the high-pressure limit (i.e.,

statistical equilibrium) For reactions that are pressure dependent, more sophisticated

methods such as RRKM28 rate calculations coupled with master equation29 calculations

19 (to estimate collisional energy transfer) allow for estimation of low-pressure rates. Rate

coefficients obtained over a range of temperatures can be used to obtain two- and three-

parameter Arrhenius expressions:

⎛ −E ⎞ ⎜ a ⎟ k(T) = Ae⎝ RT ⎠ (1.23)

⎛ −E ⎞ ⎜ o ⎟ k(T) = AT me⎝ RT ⎠ (1.24)

Common methods for exploring the energetics of elementary reaction steps

include ab initio30 and density functional theory (DFT).31 As computational speeds have

increased, use of higher levels of theory have allowed for more accurate prediction of

properties and reactions for reactive radical intermediates, further advancing our understanding of combustion chemistry.

Using these methods, the elementary reaction steps that define a fuel’s overall

combustion can be compiled, generating an overall combustion mechanism. Current

combustion software allows a user to provide a mechanism for a specific fuel and select

other parameters of interest, along with a reactor model that simulates a particular

reaction environment via a given application code (Figure 1.4).32 One such application

code can simulate a flame in this fuel, thus providing a wealth of information about a

flame’s speed, key intermediates, and dominant reactions. Computational fluid dynamics

can be combined with detailed chemical kinetic models to also be able to simulate

turbulent flames and combustion environments.

While theoretical calculations generally have been used to supplement experimental findings, they also hold enormous promise for fully discerning the potential

20 energy surfaces of relevant combustion pathways, as well as identifying and exploring the chemistry of relevant reactive intermediates.

KINETIC MECHANISM

THERMODYNAMIC REACTOR MODEL RESULTS DATA

TRANSPORT DATA

Figure 1.4: Flow chart for a typical master equation method.

1.3. Reactive Radical Intermediates in Combustion Chemistry

1.3.1. ALIPHATIC SYSTEMS

1.3.1.1. Methane combustion. The simplest hydrocarbon, methane, has posed a wealth of challenges to experimentalists and theoreticians seeking to discern its combustion mechanism. These reactions have been explored in a wide variety of contexts over the

21 past several decades. We have discussed these briefly; the interested reader is referred to

the reviews cited in our previous discussion chemistry for further details. Due to the scope of this chapter, we are primarily interested in these reactions insofar as they

• provide useful benchmarks for the reactions of larger alkylperoxy (RO2 ) and alkyloxy

(RO•) systems. With respect to the reactive intermediates present in methane combustion

and their implications for larger systems, Lightfoot has published a review on the

atmospheric role of these species,33 while Wallington et al. have provided multiple

overviews of gas-phase peroxy radical chemistry.34 Lesclaux has provided multiple

reviews of developments in peroxy radical chemistry.35 Batt published a review of the

gas-phase decomposition reactions available to the alkoxy radicals.36

Notably, the Gas Research Institute’s mechanism (GRI-MECH) for methane combustion is well-established, drawing on research from several groups over several decades to define and calibrate kinetic and thermodynamic data for each elementary reaction step. Additional mechanisms37 for methane oxidation are also available and

updated periodically to include the most recent data.

• • Methoxy (CH3O ) and methylperoxy (CH3O2 ) radicals have been subjected to

substantial study. Zaslonko et al. have reviewed several reactions involving the methoxy

radical.38 More specifically, several experimental methods have been employed in

studying this species. As early as fifty years ago, methoxy radical was an experimental

target in mass spectrometric studies by Lossing.39 Burcat and Kudchadker used IR

vibrational spectra of at varying temperatures to extrapolate the ideal gas

properties of methoxy radical.40 Ruscic and Berkowitz used photoionization mass

22 41 spectrometry to determine the ionization potential (IP) and heat of formation (ΔHf) of methoxy radical. Martinez et al. used laser-induced fluorescence (LIF) to determine the

• • temperature and pressure dependence of the rate coefficients for CH3O + NO2 over the

250 to 390 K temperature and 50 to 600 Torr pressure ranges.42 Wollenhaupt et al.

examined the same reaction using pulsed laser photolysis (PLP).43 Computationally,

Carter and Cook completed an extensive assessment on theoretical approaches for

methoxy radical,44 while Page explored the kinetics of its unimolecular decomposition.45

Subsequent studies have focused on methoxy radical’s reactivity with a variety of

46 47 • 48 • molecules: Pan et al. and Sun et al., with NO2 ; Pang et al., with NO . Gomez et

al.49 have modeled the addition and H-atom abstraction reactions of methoxy radical with

various hydrocarbons.

In early studies of methylperoxy radical, Simonaitis and Heicklin50 observed its

• • 51 reaction with NO and NO2 , while Cox and Tyndall used molecular modulation spectroscopy to supplement these findings, and Kan and Calvert studied the water vapor

• 52 dependence of the CH3O self-reaction. More recently, Wallington et al. obtained

absorption spectra of methylperoxy and other alkylperoxy radicals, over the 200-400 nm

range.53 Tyndall et al. explored the self-reaction of methylperoxy radical via Fourier-

Transform Infrared (FT-IR) spectroscopy, while Ghigo et al. have modeled the reaction

energy surface computationally.54 Their findings complemented an earlier flash

photolysis (FP) study by Lightfoot et al.55 Lesar et al. completed a quantum mechanical

• • 56 investigation of the reaction of CH3O2 with NO . Biggs et al. explored the reaction of

•57 • methylperoxy radical with NO3 as a possible source of night-time HO radical.

23 • Atkinson and Spillman have explored the kinetics of CH3O2 using cavity ring-down

spectroscopy (CRDS), confirming previous findings by Hunziker58 and Pushkarsky.59

Enami et al. explored the kinetics of the reaction between bromine monoxide (BrO•) and

methylperoxy radical.60 Together, these experimental and computational investigations provided mechanistic insights, rate constants, pressure dependences, branching ratios, etc. for the creation and validation of the 325-step GRI-MECH for methane combustion.

1.3.1.2. Ethyl radical + O2. Intramolecular reactions become increasingly important as

• the size of the alkyl chain increases. Ethylperoxy (CH3CH2O2 ) radical is a target of

significant interest because it can undergo intramolecular rearrangement via a low-strain,

5-membered ring transition state structure. Historically, ethyl radical has been implicated

as a discrete intermediate in many reactions. While early experimental studies of ethyl

radical suggested that this species reacted bimolecularly with O2 to yield ethene

• 61 (H2C=CH2, C2H4) and hydroperoxyl (HO2 ) radical, subsequent observation of other

products such as acetaldehyde and oxirane implied a more complex mechanism. In 1975,

Hickel62 affirmed that this initial view of ethyl radical oxidation was incomplete, with

support by Dechaux and Delfosse63 in 1979. Rather, it was deemed likely that ethyl

radical oxidation proceeded through a Semenov-type mechanism (Equation 1.3), in which

ethylperoxy radical is an intermediate, formed by collisional cooling, which could

transfer a hydrogen atom intramolecularly and decompose via multiple pathways.

Subsequent experimental work bolstered the likelihood that ethylperoxy radical was an

ethane oxidation intermediate. Baldwin postulated in 1986 that molecular oxygen adds to

24 ethyl radical to form ethylperoxy radical, which then undergoes concerted H-atom

transfer and elimination to yield ethene and hydroperoxyl radical.64 These and other

experimental findings have been summarized in several reviews, including those of

Fish,65 Walker,66 and Pilling et al.67

• The small size of the CH3CH2 + O2 system makes high-level computational exploration of its reaction energy surface tractable. Rienstra-Kiracofe et al. have

• provided an excellent review of the advances made in ab initio modeling of the CH3CH2

68 + O2 potential energy surface over the past several years. These authors also

summarized their own high-level calculations, noting the five most plausible pathways for the ethyl radical + O2 reaction, four of which involve the unimolecular

decomposition/rearrangement of ethylperoxy radical.

• • C2H5 + O2 → C2H4 + HO2 (1.25)

• • • C2H5 + O2 → CH3CH2OO → CH2CH2OOH

• → c-CH2CH2O + HO (1.26)

• • • • C2H5 + O2 → CH3CH2OO → CH3 CHOOH → CH3CHO + HO (1.27)

• • • • C2H5 + O2 → CH3CH2OO → CH2CH2OOH → C2H4 + HO2 (1.28)

• • • C2H5 + O2 → CH3CH2OO → C2H4 + HO2 (1.29)

Equation 1.25 accounts for the NTC range observed for the ignition of ethane.

Essentially, these reactions are refinements of the Semenov mechanism, since unimolecular reactions are important pathways in the oxidation of ethane.

25 Due to its importance to hydrocarbon combustion as a model alkylperoxy radical,

ethylperoxy radical continues to be the subject of experimental studies. Xing et al. used

time-resolved, negative-ion mass spectrometry to explore the reaction of ethylperoxy radical and .69 In separate works, Matricq and Szente explored the kinetics of

ethylperoxy’s reaction with acetylperoxy radical70 and NO71 using transient diode laser

absorption and time-resolved ultraviolet (TR-UV) spectroscopy. Atkinson and Hudgens

have used ultraviolet cavity-ringdown spectroscopy (CRDS) to perform kinetic studies on

the ethylperoxy radical self-reaction,72 while Rupper et al. used CRDS to explore the

• 73 A˜ ← X˜ electronic transition of CH3CH2O2 . Hasson et al. used FT-IR and high-

performance liquid chromatography (HPLC) with fluorescence detection to study the

• • 74 reaction of CH3CH2O2 with HO2 , while Mah et al. produced a mid-IR spectrum of this

species.75

• Ethoxy radical (CH3CH2O ) has enjoyed considerable interest as well. Choi et al.

explored its photodissociation dynamics via photofragment translational spectroscopy,76 while Faulhaber et al. pursued the same goal using photofragment coincidence imaging.77

Computationally, Matus et al. performed CCSD(T) calculations78 to generate quantities

of interest for this radical, including computing a heat of formation (–2.7 ± 0.8 kcal/mol)

via atomization energies.

The development of an ethane combustion mechanism provides a historical

context for understanding some overall trends of alkyl radical combustion. An

understanding of the likely pathways for this small system is useful in modeling

26 chemistry of larger systems, as can be observed from an examination of some other

reactive radical intermediates.

• 1.3.1.3. n-Propylperoxy radical (CH3CH2CH2O2 ). In longer chain hydrocarbon radicals, isomerization reactions become more important; these pathways compete with bimolecular oxidation reactions and can impact ignition rates at low temperatures. For instance, n-propylperoxy radical can undergo several unimolecular dissociations/rearrangements:79

• • CH3CH2CH2OO → CH3CH2CH2 + O2 (1.30)

• • CH3CH2CH2OO → CH3CH=CH2 + HO2 (1.31)

• • CH3CH2CH2OO → CH3CH2CH OOH (1.32)

• • CH3CH2CH2OO → CH3CH CH2OOH (1.33)

• • CH3CH2CH2OO → CH2CH2CH2OOH (1.34)

.

Reversion to reactants (Equation 1.30) is straightforward; similarly, the concerted H-atom

transfer and elimination pathway (Equation 1.31), observed for ethylperoxy radical can

• also occur for CH3CH2CH2O2 . However, multiple isomerization pathways (Equations

1.32, 1.33, 1.34) are now possible because the length of the alkyl chain has increased. In

particular, the 1,5-H-atom transfer occurs via a six-membered ring transition state, thereby minimizing ring strain. Rearrangement reactions lead to a more complex mixture of products. Extrapolating from our knowledge of the ethylperoxy radical pathways, these isomerization products can decompose into , propanal, and other oxidation

27 products. The preferences of n-propylperoxy radical will impact overall propane

combustion.

Because n-propylperoxy radical is small and can undergo low-barrier rearrangements, it is important to understand its possible rearrangement products; these findings have implications for larger hydrocarbons. DeSain et al. studied the production of hydroperoxyl radical via the reaction of n-propyl radical with O2 and proposed an

activation barrier of 26.0 kcal/mol for Reaction 1.31;80 moreover, they have also

observed sizable amounts of HO• radical, likely following one of the unimolecular H-

atom transfer steps.81 Zalyubovsky et al. examined the A˜ ← X˜ transition of

• 82b CH3CH2CH2O2 (using CRDS) and detected several of its rotational isomers at 298 K.

• • Chow et al. explored the kinetics of CH3CH2CH2O2 + NO via high-pressure chemical

ionization mass spectrometry (CIMS).83 Kaiser explored the production of propene from

• the CH3CH2CH2 + O2 reaction as a function of temperature and pressure and posited that

• chemically activated CH3CH2CH2O2 was an intermediate, resulting from the addition

reaction’s exothermicity.84 Computationally, DeSain et al.85 performed QCISD(T)

• studies to explore the 1,4- and 1,5-H-atom transfer reactions for several RO2 systems

• • (including CH3CH2CH2O2 ), generating RRKM/master equation rates to model HO2 and

HO• production. Naik et al.86 performed a similar study using quantum RRK rates to

yield rates for hydroperoxyl radical production in the n-propyl + O2 system. Chen and

Bozzelli have completed ab initio and DFT studies on the thermochemical and kinetic

parameters of the n-propyl + O2 reaction; they hypothesized that low-temperature ignition

• for this system results from intramolecular H-transfer reactions of CH3CH2CH2O2

28 followed by addition of a second oxygen molecule making it capable of forming products

that undergo chain-branching.87

• Our group has examined the conformations of CH3CH2CH2O2 , as well as its

possible unimolecular decomposition pathways.88 n-Propylperoxy radical conformations

are described by two torsion angles: CC–CO in a trans (t) or gauche (g) orientation and

CC–OO in a trans (T) or gauche (G) orientation. Energy profiles for these torsions were generated computationally, minima were obtained, enantiomers were noted, and five

unique conformers were obtained (Figure 1.5). These conformers could rapidly

interconvert, as the highest torsional barrier is predicted to be less than 5.0 kcal/mol.

Using the CBS-QB3 theoretical method,89 the 298 K distribution of rotamers was

calculated to be 28.1, 26.4, 19.6, 14.0, and 11.9 % for the gG, tG, gT, gG’, and tT

conformers, respectively. Therefore, all five conformers will be present as predicted experimentally by Zalyubovsky et al.82b

29

Figure 1.5: Conformers of n-propylperoxy radical, named according to the conformational preferences around the central C−C bond, then around the C−O bond. Figure courtesy of John Merle (J. Phys. Chem. A 2005, 109, 3637-3646). Reprinted with permission of J. Phys. Chem. A.

• The unimolecular reactions of CH3CH2CH2O2 were studied in detail (Figure 1.6);

complete potential energy surfaces were generated using both DFT (B3LYP/6-31+G(d,p)

and mPW1K/6-31+G(d,p))31,90 and CBS-QB3 methods. As expected, 1,5-H transfer

(Equation 1.34) occurs with the lowest barrier, followed by simultaneous 1,4-H transfer

• and HO2 expulsion (Equation 1.31). The overall decompositions for each H-atom transfer to yield decomposition products were modeled. It was shown that, although the

• 1,5-H-transfer reaction has the lowest barrier, the resultant CH2CH2CH2OOH cannot

easily undergo further unimolecular rearrangements. Rather, the 1,4-H atom transfer

routes (Equations 1.31 and 1.33) encounter lower barriers in subsequent steps. This

30 pathway for n-propylperoxy radical parallels those for the decomposition of ethylperoxy

radical.68

Figure 1.6. Initial reaction barriers for unimolecular reactions of n-propylperoxy radical. Figure courtesy of John Merle (J. Phys. Chem. A 2005, 109, 3637-3646). Reprinted with permission of J. Phys. Chem. A.

The CBS-QB3 potential energy surface accounts for the various experimentally- observed products, including hydroperoxyl radical, propene, HO•, propanal, and oxirane

• (c-C3H6O). The activation barrier for simultaneous 1,4-H transfer and HO2 expulsion,

obtained via calculations, compared well to the experimentally-observed barrier (26.0

kcal/mol) of DeSain et al.80,81 This work provides some ramifications for larger

alkylperoxy radicals: multiple conformers of long alkylperoxy radicals are likely to play a

31 role in the overall oxidation chemistry and dictate consideration for correct treatment of

thermochemistry; at lower temperatures (T < 500 K), unimolecular reactions dictate

peroxy radical chemistry.

• 1.3.1.4. n-Butoxy radical (CH3CH2CH2CH2O ). Just as n-propylperoxy radical is the

smallest peroxy radical that can undergo the 1,5-H-atom transfer, n-butoxy radical is the

smallest alkoxy radical that can do so, while pentyl radical is the smallest alkyl radical

with this potential. Similarly, these species are often used as models to understand the

chemistry of larger alkoxy radical and alkyl radical systems.

n-Butoxy radical exists in multiple conformer forms and can undergo a facile 1,5-

H-atom transfer. Vereecken and Peters91 exhaustively examined this possibility, via DFT

calculations and TST rate coefficients, to recommend a rate coefficient of 1.4 x 105 s-1, which agreed well with the experimental rates of both Atkinson92 and Hein et al.93

Vereecken and Peters used multiple approaches for deriving their multirotamer transition- state theory expressions, and demonstrated consistency through all of these. Ferenac et al. examined this unimolecular isomerization (1,5-H shift) for n-butoxy radical and its functionalized derivatives, noting substantial substituent effects (more dependent on substitution patterns than on the functional groups themselves).94 Lendvay and Viskolcz examined unimolecular reactions available to n-butoxy radical via ab initio and RRKM calculations and noted that, while 1,5-isomerization was the fastest route, fragmentation reactions would compete at combustion temperatures.95 This finding was corroborated

by exhaustive quantum chemical/RRKM dynamics calculations by Somnitz and

32 Zellner.96 Jungkamp et al. generated an exhaustive atmospheric mechanism for n-butane

via DFT and ab initio methods, proposing that n-butoxy radical will react primarily via

1,5-H transfers to ultimately form 4-hydroxy-1-butanal, while 2-butoxy radical will tend

to decompose via β-scission to ethyl radical and acetaldehyde.97 Cassanelli et al.

completed relative rate studies of 1-butoxy radical using FT-IR spectroscopy, noting that

reaction with oxygen competed with isomerization.98

• 1.3.1.5. 1-Pentyl radical (CH3CH2CH2CH2CH2 ). For pentyl radical, 1,5-H transfer can

occur regardless of whether further O2 oxidation occurs. Thus, this reaction has

implications for low-temperature combustion, as this (unimolecular) isomerization can

directly compete with the bimolecular reaction with O2. Most simply, n-pentyl radical

can quickly isomerize to iso-pentyl radical; each of these radicals can undergo β-scission reactions to yield a new alkyl radical + alkene:

• • • C3H7 + H2C=CH2  n-pentyl ' iso-pentyl

•  H2C=CHCH3 + C2H5 (1.35)

Several experimental studies, over the past several decades, have modeled the overall

combustion of this system.99 Jitariu et al. have calculated unimolecular rates for

reactions of pentyl radical (i.e., intramolecular H-atom transfer, β-scission, and

elimination) noting that isomerizations have lower barriers than β-scissions.100

1.3.1.6. Larger aliphatic species. In general, the trends predicted by n-propylperoxy

radical, 1-butoxy radical, and n-pentyl radical provide good benchmarks for

33 understanding the chemistry of longer chain radicals.101,102 For instance, the 1,5-H-atom

transfer and the 1,4-H-atom transfer concurrent with elimination provide two important

routes with implications for both low and high-temperature combustion. These

possibilities constitute an important sub-class of the reactions included in comprehensive

mechanisms103 for the corresponding parent compound (i.e. propane,104 butane,105 and pentane). Also, developing larger mechanisms by substituting reaction data for similar, smaller species is a common practice in mechanism development (referred to as lumping106); thus, the reactions of these model compounds have parallels in the

chemistry of larger fuels (in particular, n-heptane107 and iso-octane,108 which together

constitute the primary reference fuels used to model gasoline combustion,109 as well as n-

hexadecane, which shows promise for understanding diesel oxidation110). Chemistry

taking place through 5- and 6-membered ring transition states is consistently favored

kinetically over larger and smaller membered ring transition states.

Gasoline and diesel fuels are two aliphatic hydrocarbon fuels that merit further

discussion.111 These fuels are primarily used in different types of combustion

environments: gasoline, in a spark-ignition (SI) engine, and diesel, in an autoignition

engine. A spark-ignition engine relies on a four-stroke internal combustion process,

involving the reciprocating piston, the intake valve, the exhaust valve, and a spark plug.

In terms of the mechanism of combustion, a spark plug ignites the compressed fuel-air

mixture, and the resultant flame ideally propagates smoothly across the engine cylinder.

The speed at which the flame propagates is dependent on the fuel used. In a diesel

engine, ignition relies on compression of the fuel until the autoignition temperature can

34 be reached; no spark ignition is used, and no flame propagation occurs, and thus fewer emissions are involved. Similarly, autoignition temperature varies between given fuels.

Even from this general overview, it is clear that the efficiency at which either type of engine operates depends heavily on fuel identity. Common metrics for understanding the ignitability of gasoline and diesel fuel are referred to as the octane number and the cetane number, respectively. Essentially, octane number refers to the volume percent of iso-octane in a given gasoline sample; cetane number refers to the volume percent of cetane (n-hexadecane) in a given diesel sample. In practice, the octane and cetane numbers refer to the practical efficiency of a given fuel to that of iso-octane and n- hexadecane, respectively. Wallington et al. have recently provided an excellent tutorial review of these and other related topics involving chemistry’s many roles in automotive fuels and engines.

1.3.2. AROMATIC SYSTEMS

1.3.2.1. Soot formation. Before we examine the oxidation pathways available to aromatic systems, it is first instructive to review the most notorious role of these compounds in combustion chemistry: their propensity to lead to soot-formation. Soot is a byproduct of fuel-rich combustion, and soot particles can affect respiration and general health in humans.112 Soot production is a result of polycyclic aromatic hydrocarbon

(PAH) formation in flames: as reactive hydrocarbon radical intermediates combine to grow and propagate, they can also cyclize into rings, which ultimately yield large networks of aromatic rings.113 To work towards minimizing harmful soot emissions, it is

35 important to understand the mechanisms by which soot forms. Generally, the first

cyclization step, whereby a benzene ring is formed from acyclic radical species, has been

postulated to occur by one of two steps based on (C2H2) experiments: (1) the

• reaction of acetylene with either 1,3-butadien-1-yl radical (H2C=CH–CH=CH ) or buta-1-

• • en-3-yn-1-yl radical (HC≡C–CH=CH ); (2) propargyl radicals (HC≡C–CH2 ) self- reaction, followed by H-atom transfers. More recently, cyclopentadiene has also been implicated as a likely precursor to benzene formation.114

Once the initial benzene ring has cyclized, it can undergo sequences of H-atom

abstraction followed by acetylene addition, to yield polycyclic aromatic hydrocarbons

(PAHs). This is known as the H-abstraction-C2H2-addition (HACA) process, proposed

by Frenklack and Wang.115 As an aromatic species aggregates to a size over 500 amu, it

adopts a particulate form and can coalesce with other PAHs to further increase in size.

When many of these particles agglomerate, they form soot.116 Efforts to minimize soot

production are widespread. Notably, decreasing the carbon content relative to oxidizer

concentration in a fuel/oxidizer mixture decreases the amount of soot formed.

1.3.2.2. Benzene and toluene. In addition to their roles in soot formation, aromatic

compounds undergo oxidation processes unique from acyclic saturated hydrocarbons.

Benzene and toluene are common fuel additives; they reduce undesirable autoignition

events (engine knock), thereby, increasing a fuel’s octane rating. Therefore, aromatic oxidative decompositions have implications for combustion and atmospheric chemistry.

The C–C and C–H bonds present in aromatic hydrocarbons are substantially stronger than

36 those of , due to their sp2 hybridized carbons. When we consider the combustion reaction of an unsaturated species such as benzene or toluene, a new initiation step is

possible. A reactive radical (for instance, HO•) may either abstract a hydrogen atom or

add directly to the ring’s π-system, generating an allylic-type radical system: two

competing pathways (Figure 1.7).

+H2O

OH

+HO H

OH

HO H

HO H

Figure 1.7. Reactions of toluene with HO• radical. HO• can abstract a benzylic hydrogen atom or add to the aromatic ring at the ipso, ortho, meta, and para positions relative to the methyl group. Each resultant radical can decompose by various pathways, depending on temperature and pressure.

Both pathways contribute to the combustion chemistry of aromatic species; HO• addition

to the aromatic ring is the more prevalent at 298 K.117 In a mono-alkyl-substituted

37 aromatic species, such as toluene, abstraction of a hydrogen atom from the side chain can

compete with HO• addition at positions ipso, ortho, meta, or para to the side chain. Thus,

the possibilities for oxidative initiation and subsequent peroxy radical reactions increase.

1.3.2.3. Benzene oxidation. Of the aromatic hydrocarbons, the oxidative pathways of

benzene have been studied most exhaustively. Fujii et al.118 proposed a global

mechanism, in the early 1970s in which the C−H bond of benzene is broken to form the

• phenyl (C6H5 ) radical that subsequently reacts with molecular oxygen to form the

• phenylperoxy (C6H5OO ) radical:

• • • C6H5 + O2 → C6H5OO → 2CO + C2H2 + C2H5 (1.36)

Although greatly simplified, this model demonstrates the importance of phenylperoxy

• radical (C6H5OO ) as a reactive intermediate and accounts for some of the major combustion products. However, several other mechanistic intermediates (C3, C4, and C5 hydrocarbons) were also observed.

Subsequently, Glassman’s mechanism for benzene combustion accounted for

• several more products, proposing that phenoxy (C6H5O ) radical was the chief reactive

intermediate driving the combustion of benzene. The stepwise mechanism for benzene

119 • combustion began with H-atom loss to form a phenyl radical (C6H5 ), then proceeded through reaction with O2 and then carbon monoxide expulsion to form cyclopentadienyl

• (c-C5H5 ) radical, which could react with O2 and expel CO again to form smaller

hydrocarbon species (Figure 1.8). However, this model still over-predicted the formation

of phenyl, phenoxy, and cyclopentadienyl radicals; moreover, it failed to account for

38 additional experimentally observed products (i.e. furan, pyranyl radical, and oxo-

butadiene).120

O2

+HO2

O

O2 +O(3P)

O

+CO

O

O2 +O(3P)

O

+CO

+

Figure 1.8: High-temperature oxidation pathways of benzene, as proposed by Glassman.

The tendency of the benzene combustion mechanism to substantially overestimate

the formation of phenoxy radical suggested either flaws in the kinetic data involving

• C6H5O or incompleteness of the combustion mechanism. Subsequent work verified the

• kinetics and energetics used in modeling C6H5O decomposition. In separate studies, Liu

et al.121 and Olivella et al.122 used ab initio and DFT models, along with RRKM rate

39 coefficients, to confirm the data for these pathways. (More recently, these data were also

borne out in a quantum mechanical/RRKM study by Hodgson et al.123)

Consequently, the benzene oxidation mechanism was further developed by

considering additional decomposition and oxidation steps. Sethuraman et al. proposed

that phenyl radical decomposition can occur by either of two key pathways:124 β-scission

of phenyl radical or by breakdown of the phenylperoxy radical formed by oxidation of

phenyl radical (Figure 1.9). Using PM3 calculations,125 which were ultimately verified

by DFT studies,126 Carpenter predicted that another species, 2-oxepinoxy radical (3 in

Figure 1.9b), is an important intermediate due to its relative stability, formed via a spirodioxiranyl intermediate (2 in Figure 1.9b) from phenylperoxy radical. Pathway A in

Figure 1.9b is the thermodynamically preferred pathway at temperatures increasing up to

432 K, while pathway B has an entropic benefit at higher temperatures. While pathway

B essentially matched the traditional view of benzene combustion, pathway A introduced a new route for phenylperoxy radical, which could resolve discrepancies observed using previous models.

40 O O O O (a) O O O

O

PM3/UHFa 0.0 26.1 5.5 ------69.0

DFTb 0.0 27.2 22.6 11.3 46.1 -10.5 - 48.1

(b)

Figure 1.9. (a) Enthalpies (kcal/mol) leading to 2-oxepinoxy radical formation via unimolecular rearrangement of phenylperoxy radical. PM3/UHF = ΔHf and DFT (B3LYP/6-311+G(d,p)//B3LYP/6-31G(d)) = ΔH298 (b) Potential decomposition pathways for phenylperoxy radical, A involving 2-oxepinoxy radical, and B involving oxygen atom loss. Figure courtesy of Steven Kroner (J. Am. Chem. Soc. 2005, 127, 7466 – 7473). Reprinted with permission of J. Am. Chem. Soc.

41 This supposition was validated by experimental studies that demonstrated the

prevalence of different reactive oxygen species at different temperatures. Using cavity-

ringdown spectroscopy, Yu and Lin studied the reaction of phenyl radical and oxygen,

noting that phenylperoxy radical was the only adduct formed at temperatures ranging up

to 473 K;127 Venkat et al. completed flow reactor studies of benzene combustion at 1200

K and identified phenoxy radical as a key intermediate.119

Our group has completed several studies of key reactions for some relevant

reactive intermediates in benzene oxidation. Barckholtz et al. examined the oxidation

pathways of several aromatic species, using benzene as a benchmark.126a General models

were proposed for aromatic hydrocarbon and heterocycle oxidation (namely: benzene,

pyridine, furan, and thiophene) in preparation for more specific studies of each relevant

species. Carbon-centered radicals at each relevant position underwent exoergic

oxidation, and the resulting peroxy radical unimolecular decomposition pathways were

delineated. It was proposed that these peroxy radicals could undergo rearrangements that

make them of significant atmospheric interest.

A subsequent study examined phenylperoxy radical in greater detail. Fadden et

al.126b identified five possible unimolecular decomposition pathways for phenylperoxy

radical (Figure 1.10): via oxygen atom loss to form phenoxy radical (Figure 1.10, route

A); via a dioxiranyl radical species (Figure 1.10, route B); via a dioxetanyl radical

intermediate (Figure 1.10, route C), via a 1,3-peroxy radical species (Figure 1.10, route

D), and via a p-phenylquinone radical intermediate (Figure 1.10, route E). These routes

were examined using DFT (B3LYP/6-31G(d)) and ab initio (CASSCF) structures, with

42 high-level CAS-MP2, and UCCSD(T) single-point energies. Formation of the dioxetanyl species was predicted to be most favorable based on ΔG298 energies. This radical could

unimolecularly decompose to produce smaller species, including cyclopentadienyl and

pyranyl radicals, as well as acyclic oxygenated species, that are experimentally observed.

At higher temperatures (T > 500 K), the entropic benefit afforded via oxygen atom loss

becomes a contributing factor, and pathway A is the major decomposition route.

Moreover, route B leads directly to 2-oxepinoxy radical (3 in Figure 1.9b), which is

potentially an important intermediate in low-temperature benzene oxidation.

O

+O(3P)

O O A O O B

C O D O E H

O H O O

O H

Figure 1.10. Potential unimolecular reaction pathways for phenylperoxy radical.

43 The stablility of 2-oxepinoxy radical qualified it as a target for further theoretical

and experimental study. The calculations of Barckholtz et al.126a allowed the refinement

of a feasible energetic pathway toward 2-oxepinoxy radical; these DFT calculations

supplemented the semi-empirical work of Carpenter and also proposed a triradical

intermediate between the dioxiranyl and oxepinoxy species (Figure 1.10a).

Consequently, the unimolecular decomposition of 2-oxepinoxy radical (3 in

Figure 10b) was thoroughly modeled by Fadden et al. using DFT (B3LYP) methods.128

Gibbs free energy profiles (T = 298 to 1250 K) were generated. A wide range of decomposition pathways were examined, which could account for typical experimentally- observed products (Figure 1.11). Notably, the delineated decomposition pathway did not require the generation of cyclopentadienyl radical. Cyclopentadienone is a commonly observed product in benzene combustion, and most mechanisms presume that cyclopentadienyl radical is its most likely precursor. However, it was shown that 2- cyclopenten-1-yl radical (19 in Figure 1.11), a feasible source of cyclopentadienone, could be formed from the 2-oxepinoxy radical. These computer pathways supported the experimental results of Pfefferle and coworkers,137b in their work on the low-temperature

combustion of benzene, in which C2, C3, and C4 hydrocarbons were observed without

formation of C5 constituents.

44

Figure 1.11. Unimolecular decomposition pathways of 2-oxepinoxy radical (1). The relative free energies (298 K, kcal/mol) at the B3LYP/6-311+G(d,p)//B3LYP/6-31G(d) level are shown for each intermediate relative to 1, and each free energy of activation is relative to the reactant for that specific step. Figure courtesy of Michael Fadden (J. Phys. Chem. A, 2000, 104, 8121 -8130.) Reprinted with permission of J. Phys. Chem. A.

2-Oxepinoxy radical is highly stabilized by its unique structure; Mebel et al. determined, via PUMP3/6-31G(d)//UHF/ 6-31G(d) calculations, that it would have a

• relative enthalpy of ΔH0 = –91.8 kcal/mol as compared to phenylperoxy (C6H5O2 ) radical.129 The stability of this species suggests that it will be fairly long-lived, and thus

could undergo reactions of atmospheric interest, such as the addition of molecular

45 oxygen. Merle and Hadad studied the oxygen-initiated decomposition of 2-oxepinoxy

radical using both DFT and CBS-QB3 calculations.130 Calculations predicted that from T

= 298 to 750 K, O2 addition routes compete with unimolecular decomposition (Figure

1.12); for T > 750 K, however, the entropic penalty associated with the O2 addition step

causes the unimolecular decomposition of 2-oxepinoxy radical to be more favored.

O O O O O2 O2 O O O O A B O O

O O

O O O O +O(3P) O

O O +CO2 O O O

+O(3P) O

O H

O

3 +CO2 +O( P)

O O

+CO2 + H

Figure 1.12. Most favorable oxidative decomposition pathways for 2-oxepinoxy radical, at 298 K (path A) and 1250 K (path B).

46

The most stable O2 addition adduct below T = 1250 K is 6-peroxyoxepinone

(route A in Figure 1.12), which can cyclize to form a 1,4-peroxy intermediate which

subsequently releases CO2 to form a 5-oxopentanalyl radical. This species can cyclize

and fragment, yielding formyl radical, furan, and carbon dioxide. Above T = 1250 K, the

dissociative pathway (route B in Figure 1.12) afforded by 2-peroxyoxepinone’s loss of

oxygen atom becomes more favorable.

Even though 2-oxepinoxy radical (3 in Figure 9b) may be an important

intermediate in benzene oxidation, no confirmatory experimental evidence exists.

Recently, Kroner et al. published an experimental study of the gas-phase acidity of

131 2(3H)-oxepinone (C6H6O2), obtained via flowing-afterglow mass spectrometry. They postulated that this quantity could be used along with a thermodynamic cycle to determine the heat of formation of 2-oxepinoxy radical:

– + C6H6O2 → C6H5O2 + H ΔHacid (1.37)

– • – – C6H5O2 → C6H5O2 + e EA (C6H5O2 ) (1.38)

H+ + e– → H• –IP (H) (1.39)

• • C6H6O2 → C6H5O2 + H BDE = ΔHacid + EA − IP (1.40)

• • ΔHf (2-oxepinoxy radical, C6H5O2 ) = BDE + ΔHf (H ) − ΔHf (C6H6O2) (1.41)

A value of ΔHacid = 352 ± 2 kcal/mol was determined for equation 1.37. This

experimental evidence could ultimately be valuable in conclusively identifying 2-

oxepinoxy radical as a reactive species of interest.

47 The overall pathways of benzene oxidation and the decompositions of possible

intermediates have been well characterized via theoretical methods. Thus far, we have

discussed these species mainly in the context of their oxidation mechanisms, but

phenylperoxy and phenoxy radicals have also been investigated as individual

experimental targets.

• The chemistry of phenoxy (C6H5O ) radical has been of interest for several

decades. Benson et al. proposed the first rate coefficient for its unimolecular

decomposition,132 while Lin and Lin provided information on the Arrhenius parameters

for the reaction.133 Experimental and theoretical studies have examined the electronic states,134 molecular vibrational frequencies,135 and spin density, as well as thermal and

oxidative decomposition.136 Benzene combustion mechanisms rely heavily on the

inclusion of phenoxy radical data;118,,119 137 additionally, phenoxy radical decomposition

reactions are necessary in the mechanisms for combustion of several other species,

including propane,138 butane,139 and anisole.140 More notoriously, phenoxy radical has

recently been implicated in routes to dioxin and polychlorinated ; several

theoretical studies have been completed on these potential reactions.141 Finally, the roles

of phenoxy radical in flame chemistry,142 emissions,143 and other aspects of

combustion144 have been compiled in several reviews.

Phenylperoxy radical has similarly been a topic of experimental and theoretical

interest. Tokmakov et al.145 calculated a potential energy surface for phenyl radical and

O2 using ab initio G2(MP2) calculations. Weisman and Head-Gordon used time-

dependent density functional theory (TD-DFT) calculations to examine the effect of

48 substituents on the phenylperoxy radical UV-vis absorption spectrum.146 Lin and Mebel

used ab initio methods to study the phenoxy radical + O-atom reaction.147 Just et al. have

examined the A˜ ← X˜ electronic transition of phenylperoxy using CRDS.148 Tonokura et

al.149 used CRDS to study the visible absorption spectrum of the phenyl radical, as well as the kinetics of its reaction with O2. Krauss and Osman examined the UV absorption

• 150 spectra of vinylperoxy radical (H2C=CHO2 ) and phenylperoxy radicals.

Phenylperoxy radical, originally assumed to be a factor in low-temperature

combustion only, has actually been shown to play a substantial role in dictating the

overall combustion trends of benzene. Just as the isomerizations and eliminations of the

alkylperoxy radicals significantly affected their overall combustion pathways,

rearrangements and other intramolecular pathways available to phenylperoxy radical

similarly impact the overall progress of benzene combustion. This knowledge can be

extrapolated to more complex aromatic species.

1.3.2.4. Alkylated aromatics. Like benzene, toluene (C6H5CH3) is a common constituent

of gasoline. Much of the literature concerning toluene’s oxidation focuses on a global

mechanism for understanding its combustion. Emdee et al.151 proposed that toluene’s

combustion mechanism is most sensitive to its reaction with O2 to form benzyl radical

• • (C6H5–CH2 ) and HO2 . Dagaut et al. proposed that toluene oxidation is initiated by

benzylic H-atom abstraction by O2 to form the benzyl radical, which can unimolecularly

decompose to acetylene and cyclopentadienyl or react with an additional O2 and

unimolecularly decompose to phenyl and formyl radicals (via benzaldehyde, (C6H5–

49 C(=O)H).152 Pitz et al. generated a comprehensive mechanism for toluene combustion in

varying settings, due to its widespread use as a fuel additive.153 El Bakali et al. noted an

overall similarity between the chemistry of benzene and the chemistry of toluene, based

on their oxidation mechanisms.154 Ethylbenzene has also been the subject of mechanistic studies,155 most recently by Ergut et al.156

As mentioned previously, the low-temperature oxidation of toluene is proposed to

begin with either of two steps; if hydroxyl radical is present, HO• can abstract a benzylic

hydrogen atom or add directly to the aromatic ring. Once a radical is generated on the

aromatic ring or side chain of toluene, rapid oxidation can occur. The resulting peroxy

radical has several viable pathways available. Atkinson compiled data on low-

• temperature atmospheric reactions for both the benzyl radical and the C6H5–CH3/HO adduct. Andino et al. modeled the atmospheric oxidation of toluene and disubstituted xylenes, noting that several cyclized peroxy radical rearrangement products were energetically stable.157

• Additionally, studies have focused on benzylperoxy radical (C6H5CH2O2 ), although this is a less common target than phenylperoxy radical because HO• addition to

the aromatic ring is more dominant than H-atom abstraction of the benzylic C–H bond.

Elmaimouni et al. studied the equilibrium for benzylperoxy radical and benzyl radical +

O2 over the temperature range 393-433 K, extrapolating the addition reaction enthalpy to

be –20.1 kcal/mol at 298 K, and the free energy to be 11.4 kcal/mol.158 Fenter et al.

performed a kinetic study159 using the same equilibrium at 760 torr with a temperature

-13 range of 298─398 K, proposing a rate kf(T) = (7.6 ± 2.4) x 10 exp[(190 ± 160)/T(K)]

50 3 -1 -1 4 -1 cm molecule s , equilibrium Kp = 6.3 ± 0.2 x 10 atm , and reaction enthalpy ΔH298 =

–21.8 kcal/mol. Buth et al. used a flow reactor coupled to mass spectrometric detection

160 11 3 -1 -1 in a comparable study, determining kf = (4.44 ± 1.3) x 10 cm molecule s and Kp =

57,200 bar-1 at 298 K. Noziere et al. monitored the reaction of benzylperoxy radical with

hydroperoxyl radical using flash photolysis/UV absorption and continuous

photolysis/FTIR.161 El Dib et al. performed a flash photolysis kinetics study of the self-

reaction of benzylperoxy radical.162

It was again observed that rearrangement pathways comprise a substantial portion

of the oxidation routes for alkylated aromatics.10,157 Since this phenomenon is mainly

due to peroxy radical reactivity rather than of the identity of the parent compound, it is

clear that comparable rearrangements would be factors for polycyclic aromatic

hydrocarbons (PAHs), as well as for N-, O-, and S-containing heteroaromatic rings and

their alkylated derivatives.

1.3.2.5. Heteroaromatic combustion. Considering additional functionalities in an

aromatic ring allows for conclusions with implications for coal chemistry. Coal is a vital

fossil fuel; about 50% of the US is dependent on coal for electric power generation, and its use accounts for 90% percent of Ohio’s electrical power. Current engineering efforts are underway to maximize coal’s energy potential while minimizing harmful

163 environmental emissions (i.e., Hg, SOx, NOx, and CO2).

Unlike hydrocarbon-based fuels like methane and gasoline, coal has never been

subjected to a comprehensive mechanistic analysis due to the complexity of its molecular

51 structure. However, the fact that coal’s complex structure consists of monocyclic units

can be exploited: aromatic hydrocarbons and heteroaromatic rings are recurring units in

coal’s structure, even while the overall structure varies geographically. Understanding

low and high-temperature oxidation reactions for these subunits and their reactive radical

intermediates will facilitate a better understanding of their chemistry in combustion.

Heteroaromatic compounds (Figure 1.13) have been used as models for

understanding coal chemistry in several pyrolytic studies.

H N O S O 5 2 2 2 2 4 3 3 3 N Pyrrole Furan Thiophene Oxazole N N N N 2 N 2 2

3 3 5 N N 4 4 4 Pyridine Pyridazine Pyrazine

Figure 1.13. Heteroaromatic compounds of interest in coal combustion.

The azabenzenes (N-containing heteroaromatics) have historically been used most often;

their pyrolysis has been shown to proceed through reactive radical intermediates,

although the identity of these intermediates can vary with reaction conditions.164 Using a shock tube, Mackie et al. observed the pyrolytic decomposition of pyridine,165 noting that

52 three initial radicals derived from C−H bond scission: o-pyridinyl, m-pyridinyl, and p-

pyridinyl. Of these, o-pyridinyl radical was dominant and was observed to yield

(N≡C–C≡CH) through a ring-opening process, while m- and p- pyridinyl radicals formed HCN as a major product, via a less discernable pathway.

Similarly, Kiefer et al. studied the pyrolysis of pyrazine, pyrimidine, and pyridine;166 all

were observed to undergo ring-scission to yield 2-cyanovinyl radical (•CH=CH–C≡N),

which accounts for several combustion products upon decomposition, including HCN,

N≡C–C≡CH, and acetylene. 2-Cyanovinyl radical has been identified in several

pyrolytic studies of the azabenzenes, originally via the shock-tube study of Doughty et

al.164f,g These authors proposed that the relative position of the nitrogen atoms in the ring

substantially impacts the reactivity of that ring toward pyrolysis: i.e., pyrazine dissociates

more quickly than pyrimidine. Overall, the azabenzenes demonstrate increased reactivity

relative to benzene; their C–H BDEs range from 93 to 98 kcal/mol, compared to

benzene’s C–H BDE of 112 kcal/mol.167 The smallest C–H BDEs occur at positions

ortho to nitrogen (although being a C–H bond that is twice ortho to N (as in pyrimidine)

does not render a further lowering of the BDE). Kikuchi et al,168 Mackie et al.,169 and

Jones et al.170 have attributed this phenomenon to the nitrogen atom’s in-plane lone pair

electrons interacting with the lone pair electron on the carbon center of the radical,

thereby reducing the strength of the C-H bond via radical product stabilization.

The oxidative decomposition of the azabenzenes has not been studied in such

great detail. Tabares et al. studied the reaction of pyridine with O-atom, noting a

decrease in reactivity relative to benzene. 171 Alfassi et al. studied the formation and

53 reactivity of pyridylperoxy radicals in solution.172 Eisele postulated that the presence of

ions derived from pyridine and picoline in the troposphere implicates these species as atmospherically significant.173 Yeung and Elrod explored this claim via chemical

ionization mass spectrometry to study the reactions of HO• with pyridine, the picolines, the lutidines, and the ethylpyridines, and postulated that pyridinated compounds could indeed have substantial implications on tropospheric ion content.174 As with toluene, the

reaction of HO• with pyridine and alkylated pyridines is likely to proceed either via HO• addition to the aromatic ring or hydrogen-atom abstraction from a C–H bond. The new radicals can unimolecularly decompose or undergo reaction with O2. These reactions

have been modeled for toluene and the hydrocarbon analogues by Andino et al.157

Heteroaromatic compounds have the potential to add O2 at the ring nitrogen and thus

175 176 form NOx species, leading to excess and acid rain.

The five-membered heteroaromatic (furan, oxazole, pyrrole, and thiophene) are of

additional interest. Besides their role in coal combustion, these have been implicated as

emissions of biomass burning,177 residential fires,178 waste tire burning,179 cigarette smoking,180 and motor vehicles.181 Bruinsma et al. examined the pyrolysis of

heteroaromatic rings most commonly found in coal volatiles,182 determining a rank of

increasing pyrolytic reactivity: thiophene < benzene < pyridine < pyrrole <

cyclopentadiene < furan; they also noted that an additional, fused aromatic ring has a

stabilizing effect, especially for pyridine and furan. Cullis and Norris also studied the

pyrolytic processes of the heteroaromatics, yielding methane and benzene as major

products, via hydrocarbon radical intermediates. Qualitative product analysis revealed

54 that these heteroaromatics decomposed by similar mechanisms regardless of identity; the

heteroatoms were generally not present as major combustion products; they were lost as

water, hydrogen , or .183 Braslavsky and Heicklen extensively

reviewed the thermal and photochemical decomposition of heteroaromatic compounds.184

Klein et al. studied variations in heteroaromatic C–H BDEs for substituted aromatic compounds.185

In particular, furan and pyrrole, as well as their methylated derivatives, have been

common targets of pyrolytic studies. Pyrolysis of furan and its methylated derivatives

has been shown to lead to products, including carbon monoxide, acetylene, acetaldehyde,

, and allene. Grela et al. observed that methylated furan is likely to undergo C–O

• bond scission, yielding either benzene and water, via the biradical C(CH3)=CH–

• 186 CH=C(CH3)O , or isomerizing prior to decomposition to produce CO and C5H8.

Organ and Mackie also suggested that the biradical intermediate was the most likely intermediate.187 The mechanism has since been re-evaluated. Fulle et al. noted that the

major decomposition products of unsubstituted furan were formed via one of two

188 pathways, one which resulted in C2H2 and (H2C=C=O), and one which led to propyne and carbon monoxide; Sendt et al. confirmed these pathways and proposed that they were achieved via 1,2-H transfer in the original furan molecule, which lead to cyclic intermediates.189

The pyrolysis of pyrrole produces a variety of products: hydrogen cyanide,

propyne, allene, acetylene, cis-crotonitrile, and allyl cyanide, among them. Lifshitz et al.

hypothesized that pyrrole undergoes 1,2-bond (N–C) cleavage, then an internal H-atom

55 transfer to yield a radical intermediate that can isomerize to either cis-crotonitrile or allyl

cyanide, or dissociate to HCN and propyne.190 Bacskay et al. completed quantum

chemical comparisons of the isoelectronic pyrrolyl and cyclopentadienyl radicals; they

hypothesized that pyrrolyl radical is formed via C–H bond scission in the intermediate

pyrrolenine (2H-pyrrole) rather than directly via N–H bond cleavage (Figure 1.14).191

Mackie et al. explained a similar finding, postulating that it was the formation of pyrrolenine that dictated the rate at which pyrrole pyrolysis occurred.192

H N H H N N CH2

Figure 1.14. Likely pyrolysis pathway of pyrrole, via intermediate pyrrolenine.

While most studies have focused on the pyrolytic unimolecular decomposition of

these mono-heteroaromatic compounds, our group has explored their oxidative

decomposition. As with benzene, where phenylperoxy radical plays a major role in

dictating oxidation pathways, we hypothesize that the peroxy radicals derived from

heteroaromatic rings are reactive species of considerable interest for combustion and

atmospheric reactions.

Barckholtz et al. surveyed a variety of computational methods and basis sets to

select an appropriate theoretical model for study of these molecules, finding that density

56 functional theory (DFT) provided a good balance of accuracy and computational

economy.193 This study also validated the use of heteroaromatic monocyclic rings as

constructive models for their polycyclic analogues (by extension, this finding could also

confirm the similarities between the chemistry of coal and these smaller model

compounds). Bond dissociation enthalpies were compiled for a variety of poly-

heteroaromatic rings; the resulting values were compared to the corresponding mono-

heteroaromatics. For example, for benzofuran, the C2–H BDE is 117.8 kcal/mol, which

is 0.6 kcal/mol less than the corresponding C2–H BDE in furan. These calculations

showed that increasing the number of rings in the compound did not have a substantial

effect on BDE, except for a C–H bond adjacent to a bridgehead junction and for which an

electronegative heteroatom was present on the other side of the bridgehead. Even in such

cases, the deviation between the monocyclic analogue and the polycyclic derivative was

only ~ 2 kcal/mol.

In separate DFT studies, Fadden et al. examined the rearrangement pathways

(Figure 1.15) of peroxy radicals from azabenzenes194 and five-membered heteroaromatic rings.195 It was observed that each azaphenylperoxy radical can lose molecular oxygen

(1.2 → 1.1), rearrange to a dioxiranyl species (1.2 → 1.3) or a dioxetanyl species (1.2 →

1.4), or lose atomic oxygen (1.2 → 1.5). Other unimolecular decomposition pathways

afforded to alkylperoxy radicals (i.e. H-atom transfer and β-scission) are not possible for their aromatic analogs.

57 O N

O

1.3

N O

O

N O N 1.4a H O +O2

O

1.1 1.2 N O

1.4b

N O

+O(3P)

1.5

Figure 1.15. Unimolecular pathways available to heteroaromatic peroxy radicals; example shown for 2-pyridinylperoxy radical.

From the calculated energies, several main conclusions were drawn. Loss of O2 is less endoergic than loss of O-atom for most heteroaromatic peroxy radicals; two exceptions are 3-pyridinylperoxy radical, which mimics phenylperoxy radical in this aspect of its reactivity as well as many others, and 3-pyridazinylperoxy radical, in which the oxy

58 radical is a σ-radical (2A') and maintains aromaticity not observed in the peroxy radical.

Dioxiranyl formation is generally less endoergic than O2 and O-atom loss at 298 K;

however, at temperatures greater than 500 K, the entropic contributions reverse the stabilities. Dioxetanyl intermediates are more strained and, therefore, unstable intermediates; a few exceptions are observed for compounds in which an alkyl chain is adjacent to a ring nitrogen and cyclization can form stable nitrosyl radicals (as for

Pathway 1.4b for 2-pyridinylperoxy radical), but these reactions still incur high reaction barriers.

Similar routes are available to peroxy radicals of O-, S-, N-, and O,N-containing

five-membered ring heteroaromatics. The kinetic and thermodynamic parameters for the

viable arylperoxy radical unimolecular dissociation steps were calculated for the

furanylperoxy radicals; the effect of a second heteroatom was examined by a comparable

approach for the oxazolylperoxy radicals. Thermodynamic parameters for these reactions

were compared to those of the pyrrolylperoxy and thiophenylperoxy radicals. For

smaller heteroaromatic peroxy radicals, loss of O-atom to form the corresponding aryloxy

radical is preferred at 298 K to other decomposition routes. Dioxiranyl formation (Figure

1.15, Pathway 1.3) competes thermodynamically with oxygen atom loss (Figure 1.15,

Pathway 1.5) in some cases and is universally more favorable than O2 loss (Figure 1.15,

Pathway 1.1). The dioxetanyl route (Figure 1.15, Pathway 1.4) is disfavored for 5- membered ring heteroaromatic peroxy radicals, often due to formation of an anti-Bredt

double bond in the ring system. As with the azabenzenes, the dissociative pathways

became more favorable than rearrangements at high temperatures.

59 Overall, reactivity of the heteroaromatic peroxy radicals was shown to depend

heavily on ring size. For azabenzylperoxy radicals, losing an oxygen atom is

substantially unfavorable, and reversion to reactants (aryl radical + O2) is a more likely

dissociation pathway; the peroxy radical of the five-membered ring heteroaromatics can

lose oxygen atom at a lower cost. For both sets of arylperoxy radicals, isomerization

pathways are important at low temperatures. Additionally, intramolecular cyclizations

compete with O-atom loss, and some cyclizations lead to nitroso radicals, creating

implications via possible NOx formation. As with phenylperoxy radical, the

azabenzylperoxy radicals have lower barriers for rearrangement than for loss of oxygen

atom, and consequent products will influence overall combustion pathways.

1.3.2.6. Alkylated heteroaromatics. Substantially fewer studies have been published for the reactions of alkyl-substituted heteroaromatics, though these species demonstrate

comparable promise for elucidating coal combustion chemistry. Several references

discussed in the previous section contain information on methylated heteroaromatic rings.

Mackie and coworkers completed experimental196,197 and theoretical198 studies of the pyrolytic decomposition of 2-picoline (2-methylpyridine). They concluded that decomposition proceeded mainly through o-pyridinyl and 2-picolinyl radicals. The former tended to decompose predominantly to yield cyanoacetylene, while the latter favored decomposition to a cyano-functionalized cyclopentadiene (Figure 1.16).

60 CN

Figure 1.16. Cyanocyclopentadiene.

Fifteen years later, this pyrolytic study remains the most comprehensive for an

alkyl-substituted heteroaromatic compound, despite the authors’ hypothesis that the

alkylated derivatives actually provided a better model for fuel-bound nitrogen (FBN) than

did unsubstituted heteroaromatic rings, which have been more common targets.193-199

Kinetic studies are more common in the literature: Frerichs et al. examined the reaction of the picolines with oxygen atom,199 while Yeung and Elrod studied reactions of HO• with pyridine and its methyl- and ethyl-substituted derivatives. Both groups noted that the presence of nitrogen did not demonstrably affect the species’ chemistry; generally, reactivity is comparable to toluene.

The oxidation pathways for alkylated heteroaromatics start with the formation of

a radical species, via hydrogen-atom or alkyl-group homolytic bond cleavage. We

calculated these BDEs for methyl and ethyl-substituted derivatives of several key

heteroaromatics (Tables 1.1─1.3).200 Few of these experimental values exist;201 therefore, we also briefly examined the chemistry of benzylperoxy radical, because it is a hydrocarbon analogue for methylated heteroaromatics and an experimental target.

Calculations at the CBS-QB3 level closely replicated toluene’s experimentally- determined geometry, spectroscopic information, and bond dissociation enthalpy;

61 additionally, DFT (B3LYP) calculations replicated the qualitative trends predicted by the

CBS-QB3 calculations – quantitatively, the DFT calculations consistently underpredicted the BDEs and reaction energies relative to CBS-QB3.

62

X CH3 X CH2

+ H

ΔH298 ΔG298 (α - β) Methyl B3LYP CBS Experimental B3LYP CBS BDE Toluene 1 86.7 90.6 88.0-90.3b 79.4 83.8 0.72 Pyrrole 2 83.1 86.1 75.3 78.9 0.63 3 86.8 90.1 78.9 82.4 0.74 Furan 2 83.1 86.3 75.3 78.6 0.60 3 87.4 90.5 79.5 82.7 0.73 Thiophene 2 85.2 86.5 84.2 79.1 0.60 3 87.0 89.9 88.1 82.2 0.71 Oxazole 2 86.4 89.9 78.7 82.3 0.64 4 88.1 91.1 80.2 83.3 0.63 5 84.5 87.9 76.6 80.1 0.72 Pyridine 2 88.2 92.0 96.0c 80.5 84.7 0.73 3 87.0 91.0 79.6 83.7 0.72 4 87.9 91.6 80.6 84.6 0.74 Pyridazine 3 88.9 93.3 81.3 85.8 0.76 4 87.6 91.7 80.5 84.6 0.75 Pyrimidine 2 89.5 93.1 82.4 86.7 0.73 4 89.3 92.7 81.9 85.5 0.75 5 87.5 91.4 80.7 84.7 0.72 Pyrazine 2 88.0 92.3 80.5 85.0 0.72 b Reference 155 c Reference 170

Table 1.1. Thermodynamic and spin density information for hydrogen-atom loss reactions of methyl-substituted heteroaromatic rings. Enthalpies and energies in kcal/mol, obtained at the B3LYP/6-311+G**//B3LYP/6-31G* (designated as B3LYP) and CBS-QB3 (designated as CBS) levels. See Figure 1.13 for structures and numbering.

63

X CHCH X CH2CH3 3

+ H

ΔH298 ΔG298 (α - β) Ethyl B3LYP CBS Experimental B3LYP CBS BDE Ethylbenzene 1 83.9 88.1 85.4-86.9b 75.2 79.7 0.68 Pyrrole 2 80.8 84.6 72.2 75.9 0.61 3 83.8 87.7 75.3 78.9 0.71 Furan 2 80.1 83.9 71.6 75.5 0.57 3 84.2 87.9 75.4 79.1 0.69 Thiophene 2 79.8 83.8 71.1 75.2 0.55 3 83.5 87.1 74.4 78.1 0.67 Oxazole 2 82.4 86.5 74.1 78.4 0.61 4 84.5 88.3 76.0 79.9 0.68 5 81.5 85.6 73.0 77.1 0.60 Pyridine 2 84.2 88.1 75.2 80.0 0.67 3 82.9 85.6 74.8 77.2 0.69 4 83.7 87.2 74.5 78.7 0.69 Pyridazine 3 84.7 90.1 77.6 83.1 0.68 4 83.7 88.7 75.1 79.7 0.68 Pyrimidine 2 84.2 89.0 76.1 80.9 0.69 4 83.9 88.5 75.7 80.4 0.70 5 84.3 89.1 74.9 80.5 0.66 Pyrazine 2 83.8 90.2 75.1 81.4 0.68 b Reference 115.

Table 1.2. Thermodynamic and spin density information for hydrogen-atom loss reactions of ethyl-substituted heteroaromatic rings. Enthalpies and energies in kcal/mol, obtained at the B3LYP/6-311+G**//B3LYP/6-31G* (designated as B3LYP) and CBS- QB3 (designated as CBS) levels. See Figure 1.13 for structures and numbering.

64 Methyl ΔH298 ΔG298 Ethyl ΔH298 ΔG298 Pyrrole 2 106.1 89.3 2 101.1 87.1 3 104.3 87.7 3 99.1 85.6 Furan 2 108.8 92.1 2 103.9 89.9 3 105.7 89.1 3 100.6 86.8 Thiophene 2 103.9 87.3 2 99.1 85.3 3 100.9 84.2 3 95.6 81.6 Oxazole 2 109.8 93.3 2 104.9 91.1 4 106.9 90.3 4 102.1 88.3 5 110.8 94.1 5 105.9 92.1 Pyridine 2 94.1 77.9 2 89.3 75.5 3 98.3 82.1 3 93.4 79.4 4 97.6 81.7 4 92.7 78.9 Pyridazine 3 96.2 79.7 3 91.4 77.7 4 97.1 81.2 4 92.2 78.6 Pyrimidine 2 97.4 81.5 2 92.5 78.8 4 90.3 74.1 4 89.7 75.9 5 99.7 84.1 5 94.9 81.1 Pyrazine 2 94.9 78.5 2 87.1 72.3

Table 1.3. Thermodynamic information (kcal/mol, 298 K, B3LYP/6- 311+G**//B3LYP/6-31G*) for alkyl-group loss reactions of methyl and ethyl- substituted heteroaromatic rings.

65 The reactivity of methyl- and ethyl-substituted azabenzenes was explored by

calculation of the homolytic BDEs and free energies for alkyl C–H hydrogen-atom and

alkyl side-chain fragmentation. These values were analyzed as a function of heteroatom,

ring size, side chain length, spin density, and temperature. Furthermore, the impact on the

thermodynamic values derived from the harmonic-oscillator approximation was

analyzed, by treating side-chain torsions as hindered rotors. At 298 K, loss of hydrogen-

atom to form a benzylic-like radical is roughly 10 kcal/mol more favorable than loss of

the alkyl group, regardless of ring size or heteroatom; this trend is consistently exhibited

over a wide temperature range (T = 298 – 2000 K), although both reactions become

increasingly favorable with increasing temperature, due to entropic effects on the free energy (Figure 1.17). Spin densities for heteroaromatic radicals correlate well with the

BDE values: the more diffuse the spin density, the lower the corresponding BDE. This is most dramatic for the five-membered ring heteroaromatics. Both the harmonic oscillator and hindered rotor treatments give comparable values for reaction enthalpies and free energies. The ethyl derivatives have lower reaction enthalpies and free energies than methyl derivatives.

66

(a) (b)

Figure 1.17. Variation of reaction free energy (kcal/mol) with temperature (K) for alkyl C–H hydrogen-atom loss in (a) five-membered ring methyl-substituted heteroaromatic rings and (b) six-membered ring methyl-substituted heteroaromatic rings.

With respect to predicting the chemistry of larger heteroaromatic systems (such as those in coal), these calculations suggest that both hydrogen-atom and alkyl-group loss can contribute in initiation reactions to the combustion of coal; the subsequent oxidation pathways of both the aromatic peroxy radicals (previously explored194,195) and the

alkylated aromatic peroxy radicals are of interest. Reactivity will likely increase with

increasing alkylation of a subunit, and the azabenzene units are more likely to react than

the five-membered heteroaromatic rings. The initial steps of radical formation are

67 expected to become more favorable at higher temperatures, primarily due to entropic

considerations.

1.3.2.7. Oxidative decomposition of alkylated heteroaromatics. Upon formation of

alkylated heteroaromatic radicals, these species can rapidly react with O2 to form various

peroxy radicals. Again, it seems likely that these resultant reactive species will have an

impact on overall combustion processes. Benzylperoxy radical was initially explored

computationally to obtain a qualitative picture of peroxy radical decomposition for

species containing both alkyl and aromatic components; these findings were then

extended to peroxy radicals of methyl- and ethyl-substituted azabenzylperoxy radicals.202

The alkyl chains of these species are large enough than conformeric considerations are a concern. We performed calculations to analyze rotational profiles for alkyl group torsions to isolate the most stable conformation. Furthermore, in all cases, rotation of the dihedral angle occurred with a barrier less than ~5 kcal/mol. As shown in Figure 1.18, alkylated heteroaromatic peroxy radicals (1.2) can revert to reactants (1.1), or cyclize via attack of the side chain’s terminal O at a position ipso (1.3) or ortho (1.4) to the alkylperoxy chain. Oxygen atom loss can occur (1.5). Given the presence of the exocyclic group, isomerization via H-atom transfer (1.6) can occur from a methylene carbon to the terminal peroxy oxygen; this cannot occur in the non-alkylated parent compounds. Additionally, entire side chain loss (as dioxirane in the case of methyl substituted aromatics) is possible (1.7).

68 O N O N 1.3 O 1.4a

O H

OO O

N N CH 2OO N N CH2O +O2 1.4b 1.2 1.1

N CH2O +O(3P)

N O + O 1.5

N CHOOH N CHO 1.7 + HO

1.6

Figure 1.18. Unimolecular pathways for 2-picolinylperoxy radical, a representative peroxy radical for alkylated azabenzenes.

Calculated kinetic and thermodynamic values for unimolecular decomposition of benzylperoxy radical and 2-picolinylperoxy radical (2-methylpyridinylperoxy radical) at

298 K are presented in Table 1.4 (representative structures shown in Figure 1.18). The energies for decomposition of these two compounds are similar: bicyclic ring formation at the ortho carbon (pathway 1.4) occurs with a barrier of ~30 kcal/mol; ipso cyclization

(pathway 1.3) occurs with a barrier of ~35–40 kcal/mol; internal H-atom transfer

(pathway 1.6) occurs with a barrier of ~40 kcal/mol. The added functionality afforded by

69 the nitrogen atom in 2-picolinylperoxy radical allows a cyclization resulting in a stable

N–O radical after O–O bond scission (pathway 1.4b), which has potential for yielding

precursors to NOx chemistry; however, this process has a substantially high barrier (~50 kcal/mol) and is unlikely to be a factor at atmospheric temperatures. While the

qualitative trends were replicated between benzylperoxy radical and 2-picolinylperoxy

radical, the latter has slightly higher reaction barriers and energies in nearly every case,

which can be attributed to the relative stabilization of the allylic systems in the cyclized derivatives. CBS-QB3 calculations helped demonstrate that the DFT (B3LYP) approach provides good qualitative predictions for the energetic trends. The trends for 2- picolinylperoxy radical were similar for other picolinylperoxy radicals, as well as peroxy radicals of alkylated diazabenzenes (i.e., pyridazine, pyrimidine, and pyrazine). It was observed that the presence of a second ring nitrogen has little effect on either the identity

or energetics of the preferred pathways. Likewise, the ethyl analogs favored similar

reaction pathways, with energies of activation and reaction varying by only ~ 2 kcal/mol

between the methyl and ethyl analogues. When present, the disparities in the reaction

energetics were rationalized via consideration of the inductive effects caused by the

nitrogen atom(s) and varying amounts of geometric strain introduced in the

rearrangement pathways.

70

Benzylperoxy radical B3LYP CBS-QB3 Experiment ο ο ο ο ο ο ΔΗ ΔG ΔΗ ΔG ΔΗ ΔG 1.2 → 1.1 -16 -5.5 -22.7 -12.2 -20.1a, -21.8b -12.2

TS(1.2-1.3) 35.7 37.2 30.1 31.7 1.2 → 1.3 33.2 34.2 25.8 26.9 TS(1.2-1.4a) 33.5 35.6 29.7 31.8 1.2 → 1.4a 35.6 23.7 14.6 15.1 TS(1.2-1.6) 37.9 38.9 38.4 39.2 1.2 → 1.5 56.5 47.3 61.9 52.5 a a 1.2 → 1.6 -33.9 -35.4 1.2 → 1.7 53.7 42.9 55.8 45.3 a Geometry could not be optimized.

2-Picolinylperoxy radical B3LYP CBS-QB3 ΔΗ ΔG ΔΗ ΔG 1.2 → 1.1 16.3 6.5 22.7 13.2 TS(1.2-1.3) 37.1 39.4 31.0 33.4 1.2 → 1.3 34.2 35.7 26.1 27.9 TS(1.2-1.4a) 30.4 33.2 25.6 28.7 1.2 → 1.4a 20.6 23.2 13.2 16.0 TS(1.2-1.4b) 47.1 49.8 1.2 → 1.4b -5.8 -4.1 1.2 → 1.5 61.6 53.0 TS(1.2-1.6) 38.0 39.4 37.6 39.2 1.2 → 1.6 -24.7 -33.6 -23.7 -32.4 1.2 → 1.7 49.2 38.7

Table 1.4. Comparison of reaction pathway energetics (kcal/mol) for benzylperoxy radical and 2-picolinylperoxy radical at 298 K, via B3LYP/6-311+G**//B3LYP/6-31G*. All enthalpies and free energies are relative to the peroxy radical (1.2); when preceded by TS, the relative data are the enthalpy and free energies of activation. Where applicable, 1.4a refers to cyclization at a ring carbon and 1.4b refers to ortho cyclization at nitrogen.

71

Figure 1.19 shows that with increasing temperature, formation of 1.1, 1.5, and 1.7

benefit entropically and become substantially exoergic reactions. Formation of 1.3 and

1.4a exhibit little entropic benefit. As shown in Figure 1.20, the reaction barriers also gain little entropically. Formation of 1.4b has a similarly small entropic benefit but maintains a high reaction barrier, such that any NOx formation via this process is

expected to be minimal except at extremely high temperatures. Formation of 1.6 exhibits

a precipitous drop in its barrier and is the dominant processes at T ≈ 1250 K. This last

reaction is of considerable interest given its relevance to both low-temperature and high-

temperature combustion, since it essentially constitutes a dissociative rearrangement

pathway. Both the direct and indirect decomposition intermediates (1.1, 1.5, 1.6, and 1.7)

will play a substantial role in combustion as temperatures rise above 298 K, as peroxy

radicals, themselves reactive oxygen species, demonstrate the potential to generate

reactive O-atom and hydroxyl radical.

72

Figure 1.19. ΔGrxn versus temperature for the unimolecular pathways of 2- picolinylperoxy radical. 1.2 → 1.3 denoted by open diamond; 1.2 → 1.4a denoted by solid square; 1.2 → 1.4b denoted by open triangle; 1.2 → 1.5 denoted by dash; 1.2 → 1.6 denoted by x; 1.2 → 1.7 denoted by solid diamond. All energies calculated at the B3LYP/6-311+G(d,p)//B3LYP/6-31G(d) level of theory.

73

Figure 1.20. ΔGactivation versus temperature for unimolecular pathways for 2- picolinylperoxy radical. Activation energy for 1.2 → 1.3 denoted by open diamond; activation energy for 1.2 → 1.4a denoted by solid square; activation energy for 1.2 → 1.4b denoted by open triangle; activation energy for 1.2 → 1.6 denoted by x. All energies calculated at the B3LYP/6-311+G(d,p)//B3LYP/6-31G(d) level of theory.

The overall chemistry of alkylated azabenzylperoxy radicals was consistent regardless of alkyl substitution or number of nitrogen atoms. The picolinylperoxy radicals provide excellent models for the chemistry exhibited by this larger class of species. Moreover, aromatic hydrocarbons can themselves predict several aspects of this

74 chemistry, the exception being those processes involving oxidation or rearrangement with

ring .

When this approach is extended to alkylated derivatives of the five-membered

heteroaromatic rings,203 the energetic trends could be considered in light of ring size and

heteroatom effects. In general, the same six pathways are available, such that

rearrangements and dissociations are observed to contribute to the chemistry of these

species. Many trends are consistent between both sets of alkylated heteroaromatics: increasing alkyl substitution leads to small stabilization of reaction barriers and energies.

Increasing temperature leads to a shift in preference for the dissociation pathways.

However, a difference was observed for smaller rings: while cyclizations are favored pathways at 298 K regardless of parent ring size, alkyl-substituted heteroaromatic peroxy radicals of the five-membered rings cyclize in such a manner as to generate an allylic radical system (cyclizing at a position either ipso or ortho to the side chain). On the other hand, alkylheteroaromatic peroxy radicals of the six-membered rings preferred to form five-membered rings via ortho cyclization, due to their increased stability over spirodioxetane structures (which would result from ipso cyclization). This general trend is represented in Figure 1.21.

75 N N CH2OO O

O

O O

CH2OO

Figure 1.21. Qualitative depiction of favorable cyclization pathways for representative peroxy radicals of methyl heteroaromatics (top: pyridine, and bottom: furan). Cyclization for the alkylated six-membered heteroaromatics is driven by the thermodynamic stability of the resulting ring, while cyclization for the alkylated five-membered heteroaromatics is dictated by which pathway allows the generation of a stable allylic radical system.

Heteroatom identity also has an impact on the reactions of interest. Oxygen, nitrogen, and sulfur atoms affected heteroaromatic chemistry to slightly different quantitative extents (though qualitative trends were consistent), and the effect of multiple heteroatoms (as in oxazole) led to small differences in reaction energetics. Notably, the pyrrole, oxazole, and thiophene derivatives reacted to form various species wherein their heteroatoms are oxidized, and these effects were unique to each heterocycle. Thus, data for aromatic hydrocarbon species are useful in estimating the energetics of these species, but cannot completely predict their chemistry, due to the propensity of increasing functionalization to lead to different products and pathways.

76 Relative to non-alkylated heteroaromatic peroxy radicals,194,195 we observed that

alkylated species demonstrated a greater affinity for intramolecular reactions, due to the length and flexibility of their side chains. Dissociative reactions were less favorable for alkylated derivatives, due to a reduction of aromatic character. Our studies have provided important details for the oxidative decomposition of alkyl-substituted heteroaromatic rings in particular, pathways originating with peroxy radicals along with their reactivity over a wide temperature range. We have shown that the alkylated five- membered heteroaromatics demonstrate certain unique tendencies in their reactivity.

Given the differences in reactivity between the five- and six-membered heteroaromatics, as well as between the alkylated and non-alkylated heteroaromatics, it seems likely that the chemical behavior of coal could vary somewhat depending on the abundance and nature of the cyclic subunits present in its structure. However, these discrepancies are mainly limited to the variety of products formed via combustion rather than the overall kinetics and thermodynamics of the relevant processes.

1.4. Future Challenges in Combustion Chemistry

In closing this chapter, it seems logical to highlight some of the most recent

progress in combustion chemistry. Alternate forms of energy and methods of combustion

are continually being developed, thereby continuously invoking new challenges for

experimentalists and theoreticians. Some alternatives involve new combustion processes,

as in the case of homogenous-charge compression-ignition (HCCI)204 chemistry, which

depends on auto-ignition (low-temperature) chemistry within a homogenous gas mixture

77 (unlike a spark-ignition engine). This new combustion method has the potential to

increase fuel efficiency and decrease harmful emissions.205,206 One significant issue with

HCCI is adequately controlling fuel ignition during the compression stroke of the engine; this is essentially a kinetics problem, requiring understanding of the reaction rates and mechanisms of the radical species generated by the fuel. Thus, reactions normally associated with low-temperature combustion and often treated cursorily in oxidative mechanisms can have a greater impact on combustibility.

For the purposes of this chapter, we are most interested in the implications of the

changing energy landscape for the reactive pathways involved. As the natures of the fuel used in everyday life change, the reactive radical intermediates formed in their combustion will also change. We will examine the context of two classes of relevant fuel developments.

1.4.1. Fuel additives. Oxygenated molecules have long been used to augment gasoline

formulations. For instance, since the elimination of tetraethyl lead (Pb(C2H5)4), oxygenates have been used as gasoline additives to reduce harmful carbon monoxide emissions and increase a fuel’s octane rating. The most prominent oxygenate has been methyl tert-butyl (MTBE); however, this additive has since been flagged as a potential carcinogen and odorous component that can taint water supplies.207 Moreover,

MTBE necessitates the use of the industrial side-product, isobutene, for its generation,

further competing with crude oil supplies for a value-added product.

78 As the downsides to MTBE use became more apparent, ethanol emerged as an

attractive alternative oxygenate. It did not demonstrate the negative health effects of

MTBE, and moreover, it could be produced from renewable (biomass) fuels.208

Compared to MTBE, ethanol has a higher oxygen-to-carbon ratio. Ethanol has been used primarily as a gasoline additive; however, in the years since the Energy Policy Act of

1992,209 a shift has occurred, such that current manufacturers are building cars with the capacity to run on fuel blends of 85% and 95% ethanol, along with a small amount of

gasoline – these fuels are often referred to as E85 or E95, respectively. (However, some

engine modifications are necessary for vehicles to run effectively on blended fuels with

greater than 20% ethanol.) This advance has been hampered by the general lack of

availability of ethanol at fueling stations, a shortage that is being gradually remedied.

Currently, the cost benefits of using ethanol over gasoline are substantial, but the

availability of the former is still limited.210

Ethanol itself demonstrates several drawbacks as an alternative fuel, despite its increasing availability. Most notably, ethanol absorbs water. Thus, it cannot travel

through existing gasoline pipelines, as the water could subsequently separate and freeze

during colder temperatures, and possibly bursting the pipelines; moreover, ethanol is also corrosive.211 Thus, ethanol has to be transported via other means (a fact which,

ironically, generally necessitates the use of gasoline, diesel or other fuel and results in a

higher cost that negates one monetary benefit of using ethanol as a fuel). Moreover,

ethanol evaporates relatively quickly, and so in warmer temperatures, it must be blended

carefully.212

79 Butanol possesses the chemical benefits of ethanol while avoiding its drawbacks.

Its use is considerably less temperature-sensitive: it is six times less evaporative than ethanol, and it is not corrosive, so can be shipped via existing fuel pipelines.213 In terms

of fuel benefits, due to its higher number of carbons, it has higher energy content

(110,000 Btu/gallon) than ethanol (84,000 Btu/gallon), and is comparable to gasoline

(115,000 Btu/gallon); it can also be generated more easily from biomass than can

ethanol.214 Unlike ethanol, it can be used as a direct replacement for gasoline rather than as an additive therein, since butanol’s air/fuel ratio is comparable to that of gasoline.215

Finally, butanol’s primary combustion byproduct is carbon dioxide; it avoids formation

216 of the pollutants NOx, SOx, and carbon monoxide. Thus, butanol is a potential fuel of

substantial interest.

Studies have been completed specifically on combustion processes of ethanol and

butanol, and several of the peroxy and oxy radical species have been examined.92-105

Marinov proposed an exhaustive mechanism for ethanol’s combustion.217 Cavalli et al.

examined the initial reaction of HO• radical with 1-butanol by FTIR spectroscopy,218 and

Chen et al. studied the anodic oxidation of this species.219 Recent experiments by

McEnally and Pfefferle led them to propose that butanol combustion primarily occurs via a complex fission reaction rather than H-atom abstraction.220

While we have discussed the historical background of the main gasoline

oxygenates, it is worthwhile to note that several other species have been discussed in this

context as well. Notably, it has been shown that simply adding oxygen to a given

combustion environment does not in itself achieve soot reduction; the role of the oxygen

80 within the structure of the oxygenate plays a major part. For instance, Westbrook et al.

completed a modeling study on the effect of various oxygenated hydrocarbons on soot production.221 They saw a significant reduction in the number of alkynyl radicals serving

as soot precursors in a diesel flame via the inclusion of (ROH) and ether (ROR’)

additives, but noted a lesser effect when (RCO2R’) were included. This is due to

the fact that esters can readily fragment to generate CO2, so that their oxygen atoms are

not involved in the processes that can affect soot reduction. Similarly, Sinha and

222 Thomson studied three C3 oxygenated hydrocarbons - , dimethoxy

methane (DMM), and dimethyl carbonate (DMC) − all in comparison to propane, and noting different effects for each. DMM and DMC lack C–C bonds. Thus, the concentration of the alkenyl and alkynyl radicals necessary to form benzene and larger aromatic hydrocarbons was correspondingly smaller in these species, and soot production drops. Nag et al.223 have postulated that the identity of a given oxygenate contributes to the resultant balance of CO and CO2 in the radical pool; CO has a greater propensity for

PAH reduction than does CO2. The implications for power output and efficiency are

important as well, since a significant amount (~33%) of the eventual heat is generated by

• the conversion of CO to CO2 in the combustion environment (usually mediated by HO radical). The structure of an oxygenate has notable implications for both its reactivity and its propensity for the formation of emissions.

1.4.2. Biodiesel. Having alluded briefly to biomass compounds in the previous section, as sources for ethanol and butanol, we turn now to the potential of these species as

81 fundamental alternatives to fossil fuels. In 1912, Rudolf Diesel presciently stated, “The

use of vegetable oils for engine fuels may seem insignificant today. But such oils may in

the course of time become as important as petroleum and the coal tar products of the

present time.” Biodiesel can be produced from a variety of sources, 224 including animal (on small scales), algae, and vegetable oils: the fuel itself is produced via esterification of these lipids with methanol or ethanol. In terms of the overall benefits

and sources of fuels, several reviews are currently available. As early as 1987, Schwab et

al. reported on the potential of vegetable oils for forming diesel fuels.225 More recently,

Graboski and McCormick,226 Srivastava and Prasad,227 and Lin et al.228 have reviewed

the fuel properties (emissions, engine performance, etc.) of these species. Ma and Hanna

have compiled a review of the general benefits of biodiesel chemistry.229

The emergence of biofuels as potential energy sources demands a mechanistic

approach comparable to those currently in use for hydrocarbon compounds. Biofuels are

commonly esters (RCO2R’) of large long-chain fatty acids, so their reactivity will depend on more factors than these extant mechanisms. Kulkarni and Dalai have summarized mechanistic aspects of biodiesel chemistry, with respect to waste cooking oil.230

Thermolytic decomposition can yield alkanes, , , esters, and small acids.

Oxidative decomposition leads to highly functionalized peroxy radicals, which then can add a hydrogen atom to form , ultimately decomposing to aldehydes, hydrocarbons, and acids; additionally, these peroxy radicals might dimerize or oligomerize if excess oxygen is present. Hydrolytic reactions comprise an additional possibility: triglycerides readily decompose to , monoglycerides, diglycerides,

82 and free fatty acids (FFA), in the presence of water. Zhenyi et al. have completed

thermodynamic calculations on the pyrolysis of vegetable oils, postulating that, for a

given ester, the key initiation step proceeds via breaking of the alkyl (sp3) C–O bond.231

We can summarize the different possible initiation routes for a given triglyceride, familiar from our previous discussions (Figure 1.22). Even from this basic view, it is clear that several classes of reactive radical intermediates will play a role in biofuel combustion: alkyl (sp3) and alkenyl (sp2) radicals are readily formed, as well as peroxy

radicals and functionalized derivatives of each of these original species.

O

+ OH OH

H2O O

O

-H O

O

O

heat low T O O +CH3 2

O

O O

O O + O CH2CH3

Figure 1.22. Initial reaction steps for fatty acid ester decomposition, represented by ethyl butanoate, which can decompose via thermolysis (pyrolysis), oxidation, or hydrolysis.

83

Mechanism development traditionally depends on the inclusion of data for

relevant, smaller compounds; that philosophy will apply in this case, where the reactions

of both small hydrocarbons and small esters are of interest. Mechanisms for methyl

232 butanoate (CH3CH2CH2CO2CH3) have been developed by Gail et al. and Curran et

al.;233 the latter notes the negative temperature coefficient (NTC) range demonstrated by this species, suggesting that future work may benefit by exploring comparisons to

hydrocarbon chemistry. Good and Francisco have modeled the tropospheric oxidation

234 mechanism of (HCO2CH3). Metcalfe et al. derived combustion

235 mechanisms for C5H10O2 ethyl and methyl esters, using CBS-QB3 calculations.

Recent studies by Glaude et al.,236 Schwartz et al.,237 and Sarathy et al.238 have provided

experimental information on larger esters, in an effort to model further reactions with

implications for biofuel combustion. This area is of significant interest, and it is expected

that the next few years will provide a more thorough understanding of the important species and relevant mechanisms involved in biofuel combustion.

1.5. Conclusions

In the course of this chapter, we have provided a historical context for understanding combustion chemistry as it applies to some of the most fundamental hydrocarbon compounds. We have explored the implications of these processes for larger and increasingly functionalized molecules. These species can be used to model the

84 chemistry of petroleum and coal, in addition to smaller hydrocarbon fuels. Combustion

is a complex topic, such that both theoretical and experimental methods are useful in

exploring chemistry with implications for high-temperature oxidation and low-

temperature atmospheric reactions. In particular, the master equation methods developed

over the past few decades can consider the collective chemistry of thousands of

elementary steps to predict the overall oxidation of a given fuel; the kinetics and

thermodynamics of each of these elementary steps can be generated via experiment or

computational modeling.

Reactive oxygen species and other radical intermediates dictate the oxidative

decomposition of fuels. We have noted that peroxy radical intermediates provide an

enormous amount of flexibility in the combustion of a given compound, specifically in

the unimolecular steps available to that compound. In an instructive display of the

interaction of experimental and theoretical techniques, rearrangement pathways of the

peroxy radicals have been modeled computationally and provide justification for several

unexpected products.

At the start of the twenty-first century, efforts are underway to decrease society’s

dependence on fossil fuels. It is clear that alternate energy forms will bring with them their own sets of reactive radical intermediates and revisit the important intermediates seen from smaller model compounds, as we consider future challenges in combustion

chemistry. We expect that advances in experimental techniques and computational

approaches will correspondingly be developed in the years ahead.

85 References for Chapter 1

1 Gronert, S. J. Org. Chem. 2006, 71, 1209-1219.

2 Zerefos, C.S.; Isaksen, I. S. A.; Ziomas, I. Chemistry and radiation changes in the

ozone layer. Boston: Kluwer Academic Publishers, 2000.

3 Smith, G. P.; Golden, D. M.; Frenklach, M.; Moriarty, N. W.; Eiteneer, B.; Goldenberg,

M.; Bowman, C. T.; Hanson, R. K.; Song, S.; Gardiner, W. C., Jr.; Lissianski, V. V.; Qin,

Z. http://www.me.berkeley.edu/gri_mech/.

4 Compiled by Glassman, I. In Combustion, 3rd Edition. Academic Press, San Diego,

1996.

5 Schlegel, H. B. J . Chem. Phys. 1986, 84, 4530-4534.

6 (a) Semenov, N. N. Some Problems in Chemical Kinetics and Reactivity. Princeton

University Press: Princeton, NJ 1958; (b) Semenov, N. N. Chemical Kinetics and Chain

Reactions. The Clarendon Press: Oxford, 1973.

7 Kee, R. et al. Chemically Reacting Flow: Theory and Practice. John Wiley and Sons:

Hoboken, NJ, 2003.

8 Figure courtesy of Tim Barckholtz, ExxonMobil Research and Engineering.

9 Monks, P. S. Chem. Soc. Rev. 2005, 34, 376-395.

10 Atkinson, R.; Arey, J. Chem. Rev. 2003, 103, 4605-4638.

11 Pate, C. T.; Sprung, J. L.; Pitts, Jr., J. N. Org. Mass. Spec. 2005, 11, 552-555.

T12 Finlayson-Pitts, B. J.; Pitts, J. N. Chemistry of the Upper and Lower Atmosphere:

Theory, Experiments, and Applications; Academic Press, San Diego, CA, 2000.

86

13 Saunders, S. M.; Jenkin, M. E.; Derwent, R. G.; Pilling, M. J. Atmos. Chem. 2003, 3,

161-180.

14 Jenkin, M. E.; Saunders, S. M.; Wagner, V.; Pilling, M. J. Atmos. Chem. 2003, 3, 181-

193.

15 (a) Warnatz, J.; Maas, U.; Dibble, R. W. Combustion: physical and chemical fundamentals, modeling and simulation, experiments, pollutant formation. New York:

Springer, 2001. (b) Williams, A.; Pourkashanian, M.; Jones, J.M.; Rowlands, L. J. Inst.

Energy, 1997, 70, 102-113. (c) Streit, G. E.; Spall, D. E.; Hall, J. H. Environ. Sci.

Technol. 1987, 21, 519-525.

16 Tsang, W.; Lifshitz, A. Ann. Rev. Phys. Chem. 1990, 41, 559.

17 Casavecchia, P. Rep. Prog. Phys. 2000, 63, 355.

18 (a) Chen, P.; Colson, S. D.; Chupka, W. A.; Berson, J. A. J. Phys. Chem. 1986, 90,

2319; (b) Kohn, D. W.; Clauberg, H.; Chen, P. Rev. Sci. Instrum. 1992, 63, 4003.

19 (a) Rohrs, H. W.; Wickham-Jones, T.; Ellison, G. B.; Berry, D.; Argrow, B. M. Rev.

Sci. Instrum. 1995, 66, 2430-2441. (b) Zhang, X.; Maccarone, A. T.; Nimlos, M. R.;

Kato, S.; Bierbaum, V. M.; Ellison, G. B.; Ruscic, B.; Simmonnett, A. C.; Allen, W. D.;

Schaefer, III, H. F. J. Chem. Phys. 2007, 126, 44312. (c) Zhou, D.; Nag, M.; Russel, A.

L.; Read, D.; Rohrs, H. W.; Gross, M. L.; Gaspar, P. P. Sil. Chem. 2005, 3, 117-122.

20 Raffaelli, A.; Saba, A. Mass. Spec. Rev. 2003, 22, 318-331.

21 Cool, T.; Qi, F.; McIlroy, A.; Westmoreland, P.R.; Law, M.; Poisson, L.; Peterka, D.;

Ahmed, M. Rev. Sci. Instrum. 2005, 76, 941021-941027.

22 Wolfrom, J. Faraday Discuss. 2001, 119, 1-26.

87

23 Crosley, D. R. Laser Probes for Combustion Chemistry, American Chemical Society,

ACS Symposium Series, 134, Washington, DC, 1980.

24 Eckbreth, A. C. Laser Diagnostics for Combustion Temperatures and Species, 2nd Ed.,

Gordon and Breach Publishers, Amsterdam B. V., Netherlands, 1996.

25 Westbrook, C. K.; Mizobuchi, Y.; Poinsot, T. J.; Smith, P. J.; Warnatz, J. Proc.

Combust. Inst. 2005, 30, 125-157.

26 Laidler, K. J. Chemical Kinetics; Harper & Row, Publishers, Inc.: New York, 1987

27 Wigner, E. P. Z. Phys. Chem. Abst. B 1932, 19, 203.)

28 (a) Rice, O. K.; Ramsperger, H. C. J. Am. Chem. Soc. 1928, 50, 617-620. (b) Kassel,

L. S. J. Phys. Chem. 1928, 32, 1065-1079. (c) Marcus, R. A. J. Chem. Phys. 1952, 20,

359-364. (d) Gilbert, R. G.; Smith, S. C. Theory of Unimolecular and Recombination

Reactions; Blackwell Scientific Publications: Oxford, 1990. (e) Forst, W. Theory of

Unimolecular Reactions; Academic Press: New York, 1973. (f) Robinson, P. J.;

Holbrook, K. A. Unimolecular Reactions; Wiley: New York, 1972.

29 Miller, J. A.; Klippenstein, S. J. J. Phys. Chem. A, 2006, 110, 10528 -10544.

30 Hehre, W. J.; Radom, L.; Schleyer, P. v. R.; Pople, J. A. Ab Initio Molecular Orbital

Theory; John Wiley and Sons: New York, 1986.

31 (a) Parr, R. G.; Yang, W. Density Functional Theory in Atoms and Molecules; Oxford

University Press: New York, 1989. (b) Labanowski, J. W.; Andzelm, J. Density

Functional Methods in Chemistry; Springer: New York, 1991. (c) Becke, A. D. Phys.

Rev. A. 1988, 38, 3098-3100. (d) Lee, C.; Yang., W.; Parr, R. G. Phys. Rev. B. 1988, 37,

785-789. (e) Becke, A. D. J. Chem. Phys. 1993, 98, 1372-1377.

88

32 Kee, R. J.; Rupley, F. M.; Meeks, L. M.; Miller, J. A. A Fortran Chemical Kinetics

Package for the Analysis of Gas-Phase Chemical Kinetics; SAND96-8216; Sandia

National Laboratories: Livermore, CA, 1996. CHEMKIN Collection III, Reaction Design,

Inc.: San Diego, CA, 1999.

33 Lightfoot, P. D. Atmos. Environ. 1992, 26A, 180

34 (a) Wallington, T. J.; Dagaut, R.; Kurylo, M. J. Chem. Rev. 1992, 92, 667-710; (b)

Wallington, T. J.; Nielsen, O. J.; Sehested, J. “Reactions of Organic Peroxy Radicals in

the Gas Phase,” in Peroxyl Radicals, Ed. Alfassi, Z. John Wiley and Sons: New York,

1997.

35 (a) Lesclaux, R.; Lightfoot, P.D.; Veyret, B. Poll. Atmos. 1991, 10, 69-79. (b)

Lesclaux, R. NATO Science Series, IV: Earth and Environmental Sciences 2002, 16, 79-

84.

36 Batt, L. Int. J. Chem. Kinet. 1979, 11, 977-993.

37 (a) Hughes, K. J.; Turányi,, T.; Clague, A.; Pilling, M. J. Int. J. Chem. Kinet., 2001,

33, 513-538. (b) Konnov, A.A., 28th Symposium (Int.) on Combustion, Edinburgh, 2000.

Abstr. Symp. Pap. p. 317. (c) Warnatz, J. “Rate coefficients in C/H/O system.” Ch. 5 in

Combustion Chemistry, W. C. Gardiner, Jr., Ed., Springer-Verlag, New York, 1984.

38 Zaslonko, I. S.; Mozzhukhin, E. V.; Smirnov, V. N.; Tyurin, A. N. Kin.i Kat. 1991, 24,

131-153.

39 Lossing, F. P. Ann. New York Acad. Sci. 1957, 67, 499-517.

40 Burcat, A.; Kudchadker, S. Acta. Chim. Acad. Scient. Hung. 1979, 101, 249-258.

89

41 Ruscic, B.; Berkowitz, J. Preprints of Papers, Division of Fuel Chemistry. 1991, 36,

1571-1575. 202nd American Chemical Society National Meeting, New York City,

August 25-30, 1991.

42 Martinez, E.; Albaladejo, J.; Jiminez, E.; Notario, A.; Diaz de Mera, Y. Chem. Phys.

Lett. 2000, 329, 191-199.

43 Wollenhaupt, M.; Crowley, J. N. J. Phys. Chem. A 2000, 104, 6429-6438.

44 Carter, J. T.; Cook, D. B. J. Mol. Struct.: THEOCHEM 1991, 83, 111-122.

45 Page, M.; Lin, M. C.; He, Y.; Choudhury, T. K. J. Phys. Chem. 1989, 93, 4404-4408.

46 Pan, X. M.; Fu, Z.; Li, Z. S.; Sun, C. C.; Sun, H.; Su, Z. M.; Wang, R. S. Chem. Phys.

Lett. 2005, 409, 98-104.

47 Sun, Y.; Huang, X.-R.; Li, Z.-S.; Sun, C.-C. Gao. Xue. Hua. Xue. 2002, 23, 278-282.

48 Pang, X.; Feng, W.; Wang, Y.; Zhang, S. Wuli Huaxue Xuebao 1996, 12, 391-395.

49 Gomez, N.; Henon, E.; Bohr, F.; Devolder, P. J. Phys. Chem. A 2001, 105, 11204-

11211.

50 Simonaitis, R.; Heicklen, J. J. Phys. Chem. 1974, 78, 2417-2421.

51 Cox, R. A.; Tyndall, G. S. Chem. Phys. Lett. 1979, 65, 357-360.

52 Kan, C. S.; Calvert, J. G. Chem. Phys. Lett. 1979, 63, 111-114.

53 Wallington, T. J.; Maricq, M. M.; Ellermann, T.; Nielsen, O. J.; J. Phys. Chem. 1992,

96, 982-986.

54 Ghigo, G.; Maranzana, A.; Tonachini, G. J. Chem. Phys. 2003, 118, 10575-10583.

55 Lightfoot, P. D.; Lesclaux, R.; Veyret, B. J. Phys. Chem. 1990, 94, 700-707.

90

56 Lesar, A.; Hodoscek, M.; Drougas, E.; Kosmas, A. M. J. Phys. Chem. A 2006, 110,

7898-7903.

57 Biggs, P.; Conosa-Mas, C. E.; Frachebod, J. M.; Shallcross, D. E.; Wayne, R. P. J.

Chem. Soc., Faraday Trans. 1994, 90, 1205-1210.

58 Hunziker, H. E.; Wendt, H. R. J. Chem. Phys. 1976, 64, 3488.

59 Pushkarsky, M. B.; Zalyubovsky, S. J.; Miller, T. A. J. Chem. Phys. 2000, 112, 10695.

60 Enami, S.; Yamanaka, T.; Nakayama, T.; Hashimoto, S.; Kawasaki, M.; Shallcross, D.

E.; Nakano, Y.; Ishiwata, T. J. Phys. Chem. A 2007, 111, 3342-3348.

61 (a) Knox, J. H.; Wells, C. H. J. Trans. Faraday Soc. 1963, 59, 2786. (b) Knox, J. H.;

Wells, C. H. J. Trans. Faraday Soc. 1963, 59, 2801. (c) Knox, J. H. Combust. Flame

1965, 9, 297. (d) Baker, R. R.; Baldwin, R. R.; Walker, R. W. J. Chem. Soc., Faraday

Trans. 1 1975, 71, 756. (e) Baldwin, R. R.; Cleugh, C. J.; Walker, R. W. J. Chem. Soc.,

Faraday Trans. 1 1976, 72, 1715. (f) Batt, L.; Robinson, G. N. Int. J. Chem. Kinet. 1979,

11, 1045. (g) Baldwin, R. R.; Bennett, J. P.; Walker, R. W. J. Chem. Soc., Faraday

Trans. 1 1980, 76, 1075. (h) Baldwin, R. R.; Pickering, I. A.; Walker, R. W. J. Chem.

Soc., Faraday Trans. 1 1980, 76, 2374.

62 Hickel, B. J. Phys. Chem. 1975, 79, 1054.

63 Dechaux, J. C.; Delfosse, L. Combust. Flame 1979, 34, 169.

64 Baker, R. R.; Baldwin, R. R.; Walker, R. W. J. Chem. Soc., Faraday Trans. 1 1975, 71,

756.

65 Fish, A. Angew. Chem., Int. Ed. Engl. 1968, 7, 45.

91

66 Walker, R. W. In Major Research Topics in Combustion; Hussaini, M. Y., Kumar, A.,

Voigt, R. G., Eds.; Springer-Verlag: New York, 1992; p 277.

67 Pilling, M. J.; Robertson, S. H.; Seakins, P. W. J. Chem. Soc., Faraday Trans. 1995,

91, 4179.

68 Rienstra-Kiracofe, J. C.; Allen, W. D.; Scaefer, III, H. F. J. Phys. Chem. A 2000, 104,

9823-9840.

69 Xing, J.-H., Nagai, Y., Kusuhara, M., and Miyoshi, A. J. Phys. Chem. A 2004, 108, 47,

10458- 10463.

70 Maricq, M. M.; Szente, J. J. J. Phys. Chem. A, 2000, 104, 7239-7245.

71 Maricq, M. M.; Szente, J. J. J. Phys. Chem. 1996, 100, 12374-12379.

72 Atkinson, D. B.; Hudgens, J. W. J. Phys. Chem. A 1997, 101, 3901-3909.

73 Rupper, P.; Sharp, E. N.; Tarczay, G.; Miller, T. A. J. Phys. Chem. A 2007, 111, 832-

840.

74 Hasson, A. S.; Tyndall, G. S.; Orlando, J. J. J. Phys. Chem. A 2000, 108, 5979-5989.

75 Mah, D. A.; Cabrera, J.; Nation, H.; Ramos, M.; Sharma, S.; Nickolaisen, S. L. J. Phys.

Chem. A 2003, 107; 4354-4359.

76 Choi, H.; Bise, R. T.; Neumark, D. M. J. Phys. Chem. A. 2000, 104, 10112-10118.

77 Faulhaber, A. E.; Szpunar, D. E.; Kautzman, K. E.; Neumark, D. M. J. Phys. Chem. A.

2005, 109, 10239-10248.

78 Matus, M. H.; Nguyen, M. T.; Dixon, D. A. J. Phys. Chem. A. 2007, 111, 113-126.

79 Merle, J.; Hayes, C.; Zalyubovsky, S.; Glover, B.; Miller, T.; Hadad, C. J. Phys. Chem.

A. 2005, 109, 3637-3646.

92

80 DeSain, J.; Klippenstein, S.; Miller, J.; Taatjes, C. J. Phys. Chem. A 2001, 105, 4415-

4425.

81 DeSain, J.; Clifford, E.; Taatjes, C. J. Phys. Chem. A 2001, 105, 3205-3213.

82 (a) Tarczay, G.; Zalyubovsky, S. J.; Miller, T.A. Chem. Phys. Lett. 2005, 406, 81-89.

(b) Zalyubovsky, S. J.; Glover, B.; Miller, T. A.; Hayes, C. J.; Merle, J. K.; Hadad, C. M.

J. Phys. Chem. A 2005, 107, 1308-1315.

83 Chow, J. M.; Miller, A. M.; Elrod, M. J. J. Phys. Chem. A 2003, 107, 3040-3047.

84 Kaiser, E. W. J. Phys. Chem. A 1998, 102, 5903-5906.

85 (a) DeSain, J. D.; Klippenstein, S. J.; Miller, J. A.; Taatjes, C. A. J. Phys. Chem. A

2003, 107, 4415-4427. (b) DeSain, J. D.; Klippenstein, S. J.; Miller, J. A.; Taatjes, C. A.

J. Phys. Chem. A 2004, 108, 7127-7128.

86 Naik, C.; Carstensen, H.-H.; Dean, A. M. Proceedings of the Third Joint Meeting of the

U.S. Sections of the Combustion Institute, Chicago, IL, 2003.

87 Chen, C.-C.; Bozzelli, J. W. Fourth Joint Meeting of the U.S. Sections of the

Combustion Institute: Western States, Central States, Eastern States. Philadelphia, PA.

20-23 March 2005.

88 Merle, J. K.; Hayes, C. J.; Zalyubovsky, S. J.; Glover, B.; Miller, T. A.; Hadad, C. M.

J. Phys. Chem. A. 2005, 109, 3637-3646.

89 (a) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A., Jr. J. Chem. Phys. 1996,

104, 2598-2619; (b) Petersson, G. A.; Frisch, M. J. J. Phys. Chem. A 2000, 104, 2183-

2190.

93

90 Lynch, B. J.; Fast, P. L.; Harris, M.; Truhlar, D. G. J. Phys. Chem. A 2000, 104, 4811-

4815.

91 Vereecken, L.; Peters, J. J. Chem. Phys. 2003, 119, 5159-5170.

92 Atkinson, R. J. Phys. Chem. Ref. Data 1997, 26, 215.

93 Hein, H.; Somnitz, H.; Hoffmann, A.; Zellner, R. Proceedings of the Second Workshop

EUROTRAC-2 CMD, Karlsruhe, Germany, 1998.

94 Ferenac, M. A.; Davis, A. J.; Holloway, A. S.; Dibble, T. S. J. Phys. Chem. A 2003,

107, 63-72.

95 Lendvay, G.; Viskolcz, B.; J. Phys. Chem. A 1998, 102, 10777-10786.

96 Somnitz, H.; Zellner, R. Z. Phys. Chem. 2006, 220, 1029-1048.

97 Jungkamp, T. P. W.; Smith, J. N.; Seinfeld, J. H. J. Phys. Chem. A 1997, 101, 4392-

4401.

98 (a) Cassanelli, P.; Cox, R. A.; Orlando, J. J.; Tyndall, G. S. J. Photochem. Photobiol. A

2006, 177, 109-115. (b) Cassanelli, P.; Johnson, D.; Cox, R. A. Phys. Chem. Chem. Phys.

2005, 7, 3702-3710.

99 (a) Barker, J. R.; Ortiz, N. F. Int. J. Chem. Kinet. 2000, 33, 246. (b) Watkins, K. W. J.

Am. Chem. Soc. 1971, 93, 6355. (c) Watkins, K. W. J. Phys. Chem. 1973, 77, 2938. (d)

Endrenyi, L.; Le Roy, D. J. J. Phys. Chem. 1966, 70, 4081. (e) Frey, H. M.; (f) Walsh, R.

Chem. Rev. 1969, 69, 103-124. (g) Dearden, D. V.; Beauchamp, J. L. J. Phys. Chem.

1985, 89, 5359. (h) Tardy, D. C. Int. J. Chem. Kinet. 1974, 6, 291. (i) Yamauchi, N.;

Miyoshi, A.; Kosaka, K.; Koshi, M.; Matsui, H. J. Phys. Chem. A 1999, 103, 2723. (j)

Tsang, W.; Walker, J. A.; Manion, J. A. Symp. (Int.) Combust., [Proc] 1998, 27th (1),

94

135. (k) Tsang, W.; Bedanov, V.; Zachariah, M. R. Ber. Bunsen-Ges. Phys. Chem. 1997,

101, 491. (l) Pacansky, J.; Waltman, R. J.; Barnes, L. A. J. Phys. Chem. 1993, 97, 10694.

100 Jitariu, L. C.; Jones, L. D.; Robertson, S. H.; Pilling, M. J.; Hillier, I. H. J. Phys.

Chem. A 2003, 107, 8607-8617.

101 Ross, P. L.; Johnston, M. V. J. Phys. Chem. 1995, 99, 16507-16510.

102 Mereau, R.; Rayez, M.-T.; Caralp, F.; Rayez, J.-C. Phys. Chem. Chem. Phys. 2003, 5,

4828-4833.

103 Miller, J. A.; Pilling, M. J.; Troe, J. Proc. Combust. Inst. 2005, 30, 43-88.

104 (a) Curran, H. J.; Jayaweera, T. M.; Pitz, W. J.; Westbrook, C. K. Western States

Section of the Combustion Institute. Spring Meeting. Davis, CA, United States. 29

March 2004. (b) Qin, Z.; Lissianski, V. V.; Yang, H.; Gardiner, W. C.; Davis, S. G.;

Wang, H. Proc. Combust. Inst. 2000, 28, 1663-1669.

105 Marinov, N. M.; Pitz, W. J.; Westbrook, C. K.; Castaldi, M. J.; Senkan, S. M.; Melius,

C. F. Combust. Flame 1998, 114, 192-213.

106 Huang, H.; Fairweather, M.; Griffiths, J. F.; Tomlin, A. S.; Brad, R. B. Proc.

Combust. Inst. 2005, 30, 1309-1316.

107 Curran, H. J.; Gaffuri, P.; Pitz, W. J.; Westbrook, C. K. Combust. Flame 1998, 114,

149-177.

108 Curran, H. J.; Gaffuri, P.; Pitz, W. J.; Westbrook, C. K. Combust. Flame 2002, 129,

253-280.

109 (a) Curran, H. J.; Pitz, W. J.; Westbrook, C. K.; Callahan, C. V.; Dryer, F. L. Proc.

Combust. Inst. 1998, 27, 379-387. (b) Huang, Y.; Sung, C. J.; Eng, J. A. Combust. Flame

95

2004, 139, 239-251. (c) Huang, Y.; Sung, C. J.; Eng, J. A. Proc. Combust. Inst. 2002, 29,

759-766.

110 Fournet, R.; Battin-Leclerc, F.; Glaude, P. A.; Judenherc, B.; Warth, V.; Come, G. M.;

Scacchi, G.; Ristori, A.; Pengloan, G.; Dagaut, P.; Cathonnet, M. Int. J. Chem. Kinet.

2001, 33, 574-586.

111 Wallington, T. J.; Kaiser, E. W.; Farrell, J. T. Chem. Soc. Rev. 2006, 35, 335-347.

112 (a) Albagi, A.; Oja, H.; Dubois, L. Environ. Lett. 1974, 6, 241. (b) Pavanello, S.;

Simioli, P.; Lupi, S.; Gregorio, P.; Clinfero, E. Cancer Epidemiology, Biomarkers and

Prevention 2002, 11, 998. (c) Williams, J. A.; Martin, F. L.;Muir, G. H.; Hewer, A.;

Grover, P. L.; Phillips, D. H. Carcinogenesis 2000, 21, 1683-1689. (d) WHO (1997).

The World Health Report. World Health Organization, Geneva, Switzerland.

113 (a) Richter, H.; Howard, J. B. Phys. Chem. Chem. Phys. 2002, 4, 2038-2055. (b)

Miller, J. A.; Melius, C. F. Combust. Flame 1992, 91, 21-39. (c) D'Anna, A.; Violi, A.

Proc. Combust. Inst. 1998, 27, 425-433. (d) Bittner, J. D.; Howard, J. B. Proc. Combust.

Inst. 1981, 18, 1105. (e) Westmoreland, P. R.; Dean, A. M.; Howard, J. B.; Longwell, J.

P. J. Phys. Chem. 1989, 93, 8171.

114 (a) Melius, C. F.; Colvin, M. E.; Marinov, N. M.; Pitz, W. J.; Senkin, S. M. Proc.

Combust. Inst. 1996, 26, 685. (b) Friderichsen, A. V.; Shin, E. J.; Evans, R. J.; Nimlos,

M. R.; Dayton, D. C.; Ellison, G. B. Fuel 2001, 80, 1747-1755.

115 (a) Frenklach, M. Phys. Chem. Chem. Phys. 2002, 4, 2028. (b) Frenklach, M.; Wang,

H. Proc. Combust. Inst. 1991, 23, 1559.

116 Richter, H.; Howard, J. B. Prog. Energy Combust. Sci. 2000, 26, 565.

96

117 (a) Atkinson, R. J. Phys. Chem. Ref. Data, Monograph 1, 1989, 1. (b) Atkinson, R. J.

Phys. Chem. Ref. Data, Monograph 2, 1994, 1.

118 Asaba, T.; Fujii, N. Symp. (Int.) Combust., [Proc.] 14th, 1971, 155.

119 Venkat, C.; Brezinsky, K.; Glassman, I. Symp. (Int.) Combust., 19th 1982, 143-152.

120 Bittker, D. A. Prog. Energy Combust. Sci. 1991, 79, 49-72. (b) Zhang, H.-Y.;

McKinnon, J. T. Combust. Sci. Technol. 1995, 107, 261-300.

121 Liu, R.; Morokuma, K.; Mebel, A.M.; Lin, M.C. J. Phys. Chem. 1996, 100, 9314-

9322.

122 Olivella, S.; Sole, A.; Garcia-Raso, A. J. Phys. Chem. 1995, 99, 10549-10556.

123 Hodgson, D.; Zhang, H.-Y.; Nimlos, M. R.; McKinnon, J. T. J. Phys. Chem. A 2001,

105, 4316-4327.

124 Sethuraman, S.; Senkan, S. M.; Gutman, D. Combust. Sci. Technol. 1992, 82, 13-30.

125 Carpenter, B. K. J. Am. Chem. Soc. 1993, 115, 9806-9807.

126 (a) Barckholtz, C. A.; Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 1999, 103, 8108-

8117. (b) Fadden, M. J.; Barckholtz, C.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 3004-

3011.

127 Yu, T.; Lin, M.C. J. Am. Chem. Soc. 1994, 116, 9571-9576.

128 Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 8121-8130.

129 Mebel, A. M.; Lin, M. C. J. Am. Chem. Soc. 1994, 116, 9577-9584.

130 Merle, J. K.; Hadad, C. M. J. Phys. Chem. A 2004, 108, 8419-8433.

131 Kroner, S. M.; DeMatteo, M. P.; Hadad, C. M.; Carpenter, B. K. J. Am. Chem. Soc.

2005, 127, 7466-7473.

97

132 Colussi, A. J.; Zabel, F.; Benson, S. W. Int. J. Chem. Kin. 1977, 9, 161.

133 Lin, C.-Y.; Lin, M. C. Int. J. Chem. Kinet. 1985, 17, 1025.

134 (a) Land, E. J.; Porter, G.; Strachan. Trans. Faraday Soc. 1961, 57, 1885. (b) Roebber,

J. L. J. Chem. Phys. 1962, 37, 1974. (c) Ward, B. Spectrochim. Acta A 1968, 24, 813. (d)

Bent, D. V.; Haydon, E. J. Am. Chem. Soc. 1975, 97, 2599. (e) Schuler, R. H.; Buzzard,

G. K. Int. J. Radiat. Phys. Chem. 1976, 8, 563. (f) Pullin, D.; Andrews, L. J. Mol. Struct.

1982, 95, 181. (g) Johnston, L. J.; Mathivanan, N.; Negri, F.; Siebrand, W.; Zerbetto, F.

Can. J. Chem. 1993, 71, 1655. (h) Hinchliffe, A.; Steinbank, R. E.; Ali, M. A. Theor.

Chim. Acta 1966, 5, 95. (i) Chang, H. M.; Jaffe, H. H. Chem. Phys. Lett. 1973, 23, 146.

(j) Chang, H. M.; Jaffe, H. H.; Masmandis, C. A. J. Phys. Chem. 1975, 79, 1118.

135 (a) Frank, P.; Herzler, J.; Just, Th.; Wahl, C. Proc. 25th Intl. Symp. Combust. The

Combustion Institute: Pittsburgh, PA, 1994; p 833. (b) Beck, S. M.; Brus, L. E. J. Chem.

Phys. 1982, 76, 4700. (c) Tripathi, G. N. R.; Schuler, R. H. Chem. Phys. Lett. 1983, 98,

594. (d) Tripathi, G. N. R.; Schuler, R. H. J. Chem. Phys. 1984, 81, 113. (e) Johnson, C.

R.; Ludwig, M.; Asher, S. A. J. Am. Chem. Soc. 1986, 108, 905. (f) Tripathi, G. N. R.;

Schuler, R. H. J. Phys. Chem. 1988, 92, 5129. (g) Tripathi, G. N. R. In Time Resolved

Spectroscopy; Clark, R. J. H., Hester, R. E., Eds.; Wiley: New York, 1989; p 157.

136 (a) Colussi, A. J.; Zabel, F.; Benson, S. W. Int. J. Chem. Kinet. 1977, 9, 161. (b)

Harrison, A. G.; Honnen, L. R.; Dauben, H. J.; Lossing, F. P. J. Am. Chem. Soc. 1960,

82, 5593. (c) Schmoltner, A.-M.; Anex, D. S.; Lee, Y. T. J. Phys. Chem. 1992, 96, 1236.

(d) Lin, C.-Y.; Lin, M. C. Int. J. Chem. Kinet. 1985, 17, 1025. (e) Lin, C.-Y.; Lin, M. C.

J. Phys. Chem. 1986, 90, 425.

98

137 (a) Schoebel, A. Wissenschaftliche Berichte-Forschungszentrum Karlsruhe. 2000, i-

vii, 1-143. (b) Chai, Y.; Pfefferle, L. D. Fuel 1998, 77, 313-320. (c) Zhang, H.-Y.;

McKinnon, J. T. Combust. Sci. Technol. 1995, 107, 261-300. (d) Chai, Y.; Pfefferle, L.

D. Chem. Phys. Proc. Combust. 1995, 361-364.

138 (a) Marinov, N. M.; Castaldi, M. J.; Melius, C. F.; Tsang, W. Combust. Sci. Technol.

1998, 131, 1-6. (b) Marinov, N. M.; Castaldi, M. J.; Melius, C. F.; Tsang, W. Combust.

Sci. Technol. 1997, 128, 295-342.

139 Marinov, N. M.; Pitz, W. J.; Westbrook, C. K.; Vincitore, A. M.; Castaldi, M. J.;

Senkan, S. M.; Melius, C. F. Combust. Flame 1998, 114, 192-213.

140 Sanders, W. A. Gov. Rep. Announce. Index (U.S.) 1986, 86(15), Abstr. No. 633-265.

141 (a) Kim, D. H.; Mulholland, J. A.; Ryu, J.-Y. Proc. Combust. Inst. 2005, 30, 1245-

1253. (b) Kim, D. H.; Mulholland, J. A.; Ryu, J.-Y. Organohalogen Compd. 2004, 66,

1026-1032. (c) Kim, D. H.; Mulholland, J. A. Environ. Sci. Technol. 2005, 39, 5831-

5836. (d) Babushok, V. I.; Tsang, W. Chemosphere 2003, 51, 1023-1029. (e) Wiater-

Protas, I.; Louw, R. Eur. J. Org. Chem. 2001, 20, 3945-3952. (f) Grotheer, H.-H.; Louw,

R. Combust. Sci. Technol. 1998, 134, 45-63. (g) Huang, H.; Buekens, A. Organohalogen

Compd. 1998, 36, 131-134. (h) Grotheer, H.-H.; Louw, R. Symp. (Int.) Combust., [Proc.]

26th, 1996, 2405-2411. (i) Louw, R.; Grotheer, H.-H.; Kanters, M. J.; Wiater, I.

Organohalogen Compd., 1996, 27, 26-29. (j) Kim, D. H.; Mulholland, J. A. Environ. Sci.

Technol. 2005, 39, 5831-5836.

142 Hausmann, M.; Hebgen, P.; Homann, K. H. Symp. (Int.) Combust., [Proc.] 24th, 1992,

793-801.

99

143 Wang, Z.; Wang, J.; Richter, H.; Howard, J. B.; Carlson, J.; Levendis, Y. A. Energy

Fuels, 2003, 17, 999-1013.

144 (a) Sebbar, N.; Bockhorn, H.; Bozzelli, J. W. J. Phys. Chem. A 2005, 109, 2233-2253;

(b) Sebbar, N.; Bockhorn, H.; Bozzelli, J. W. Phys. Chem. Chem. Phys. 2002, 15, 3691-

3703; (c) Vovelle, C.; Delfau, J. L. NATO Science Series, Series C: Mathematical and

Physical Sciences 2000, 183-203.

145 Tokmakov, I. V.; Kim, G.-S.; Kislov, V. V.; Mebel, A. M.; Lin, M. C. J. Phys. Chem.

A 2005, 109, 6114-6127.

146 Weisman, J. L.; Head-Gordon, M. J. Am. Chem. Soc. 2001, 123, 11686-11694.

147 Lin, M. C.; Mebel, A. M. J. Phys. Org. Chem. 1995¸ 8¸407-420.

148 Just, G. M. P.; Sharp, E. N.; Zalyubovsky, S. J.; Miller, T. A. Chem. Phys. Lett. 2006,

417, 378-382.

149 Tonokura, K.; Norikane, Y.; Koshi, M.; Nakano, Y.; Nakamichi, S.; Goto, M.;

Hashimoto, S.; Kawasaki, M.; Andersen, M. P.; Sulbaek, M. P.; Hurley, M. D.;

Wallington, T. J. J. Phys. Chem. A 2002, 106, 5908-5917.

150 Krauss, M.; Osman, R. J. Phys. Chem. 1995, 99, 11387-11391.

151 Emdee, J. L.; Brezinsky, K.; Glassman, I. J. Phys. Chem. 1992, 96, 2151–2161.

152 Dagaut, P.; Pengloan, G.; Ristori, A. Phys. Chem. Chem. Phys. 2002, 4, 1846-1854.

153 (a) Pitz, W. J.; Seiser, R.; Bozzelli, J.W.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin-

Leclerc, F.; Seshadri, K.; Westbrook, C.K. 2nd Joint Meeting of the U.S. Section of the

Combustion Institute: Oakland, CA. March 25-28, 2001. (b) Pitz, W.J.; Seiser, R.;

Bozzelli, J.W.; Seshadri, K.; Chen, C. J.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin-

100

Leclerc, F.; Westbrook, C. K. Thirtieth Symposium (International) on Combustion. The

Combustion Institute; Chicago, IL: July 25-30, 2004.

154 El Bakali, A.; Dupont, L.; Lefort, B.; Lamoureux, N.; Pauwels, J. F.; Montero, M. J.

Phys. Chem. A 2007, 111, 3907-3921.

155 (a) Muller-Markgraf, W.; Troe, J. J. Phys. Chem. 1988, 92, 4914-4922; (b) Clark, W.

D.; Price, S. J. Can. J. Chem. 1970, 48, 1059.

156 Ergut, A.; Granata, S.; Jordan, J.; Carlson, J.; Howard, J. B.; Richter, H.; Levendis, Y.

A. Combust. Flame 2006, 144, 757-772.

157 Andino, J. M.; Smith, J. N.; Flagan, R. C.; Goddard, III, W. A.; Seinfield, J. H. J.

Phys. Chem. 1996, 100, 10967-10980.

158 Elmaimouni, L.; Minetti, R.; Sawerysyn, J. P.; Devolder, P. Int. J. Chem. Kinet. 1993,

25, 399-413.

159 Fenter, F.F.; Noziere, B.; Caralp, F.; Lesclaux, R. Int. J . Chem. Kinet. 1994, 26, 171-

189.

160 Buth, R.; Hoyermann, K.; Seeba, J. 25th Symp. (Int’l) on Combustion; The

Combustion Institute, 1994; 851-858.

161 Noziere, B.; Lesclaux, R.; Hurley, M. D.; Dearth, M. A.; Wallington, T. J. J. Phys.

Chem. 1994, 98, 2864-2873.

162 El Dib, G.; Chakir, C; Roth, E.; Brion, J.; Daumont, D. J. Phys. Chem. A 2006, 110,

7848-7857.

163 Sussingham, R. “Can Coal Be Clean?” Chemistry. The American Chemical Society.

Spring 2007.

101

164 (a) Axworthy, A. E.; Dayan, V. H.; Martin, G. B. Fuel 1987, 57, 29; (b) Houser, T. J.;

McCarville, M. E.; Biftu, T. Int. J. Chem. Kinet. 1980, 12, 555; (c) Houser, T. J.; Hull,

M.; Alway, R. M.; Biftu, T. Int. J. Chem. Kinet. 1980, 12, 569; (d) Leidreiter, H. I.;

Wagner, H. Gg. Z. Z. Phys. Chem. Neue Floge 1987, 153, 99; (e) Rawlins, W. T.;

Tanzawa, T.; Schertzer, S. P.; Kreck. R. H. Final Report to DOE, 1983, Contract No.

DEAC228OPC30392; (f) Kern, R. D.; Yong, J. N.; Kiefer, J. H.; Shah, J. N. 16th Int.

Symp. Shock Tubes Waves 1987, 43; (g) Doughty, A.; Mackie, J. C. J. Chem. Soc.,

Faraday Trans. 1994, 90, 541; (h) Doughty, A.; Mackie, J. C.; Palmer, J. M. 25th

International Symposium on Combustion; The Combustion Institute: Pittsburgh, 1994; p.

893.

165 Mackie, J.; Colket, M.; Nelson, P. J. Phys. Chem. 1990, 94, 4099-4106.

166 Kiefer, J.; Zhang, Q.; Kern, R.; Yao, J.; Jursic, B. J. Phys. Chem. A 1997, 101, 7061-

7073.

167 Carey, F. A.; Sundberg, R. J. Advanced Organic Chemistry: Part A, Structure and

Mechanism. New York: Kluwer Academic/Plenum Publishers, 2000.

168 Kikuchi, O.; Hondo, Y.; Morihashi, K.; Nakayama, M. Bull. Chem. Soc. Jpn. 1988,

61, 291.

169 Mackie, J. C.; Colkey, M. B.; Nelson, P. F. J. Phys. Chem. 1990, 94, 4099.

170 Jones, J.; Bacskay, G. B.; Mackie, J. C.; Doughty, A. J. Chem. Soc., Faraday Trans.

1995, 91, 1587.

171 Taberes, F.L.; Gonzalez-Urena, A. J. Phys. Chem. 1983, 87, 4933-4936.

172 Alfassi, Z. B.; Khaikin, G. I.; Neta, P. J. Phys. Chem. 1995, 99, 4544-4548.

102

173 Eisele, F. L. Int. J. Mass Spectrom. Ion Phys. 1983, 54, 119-126.

174 Yeung, L.Y.; Elrod, M. J. J. Phys. Chem. A 2003, 107, 4470-4477.

175 Williams, A.; Pourkashanian, M.; Jones, J.M.; Rowlands, L. J. Inst. Energy, 1997, 70,

102-113.

176 Streit, G. E.; Spall, D. E.; Hall, J. H. Environ. Sci. Technol. 1987, 21, 519-525.

177 (a) Molto, J.; Font, R.; Conesa, J.A. Energ. Fuels 2006, 20, 1951-1958. (b) Ciccioli,

P.; Brancaleoni, E.; Frattoni, M.; Cecinato, A.; Pinciarelli, L. Anal. Lett. 2001, 34, 937-

955. (c) Sidhu, S.; Graham, J.; Stribech, R. Chemosphere 2001, 42, 681-690.

178 (a) McDonald, J. D.; Zielinska, B.; Fujita, E. M.; Sagebiel, J. C.; Chow, J. C.; Watson,

J. G. Env. Sci. Technol. 2000, 34, 2080-2091. (b) Ozawa, T.; Ishiguro, Y.; Toyoda, K.;

Nishimura, M.; Sasahara, T.; Doi, T. Chem. Sensors 2004, 20, 56-57.

179 (a) Levendis, Y. A.; Atal, A.; Carlson, J. B. Combust. Sci. Technol. 1998, 134, 407-

431. (b) Lemieux, P. M.; Ryan, J. V. Air and Waste 1993, 43, 1106-1115.

180 Baker, R. R.; Bishop, L. J. J. Anal. Appl. Pyrol. 2005, 74, 145-170.

181 (a) Xie, H.; Wang, J.; Zhang, S.; Ding, Y.; Sun, C. J. Chem. Phys. 2006, 125, 124317-

124317-10. (b) van der Laarse, J. D.; van Leuven, H. C. E. Anal. Chim. Acta 1966, 34,

370-371.

182 Bruinsma, O. S.; Tromp, P. J. J.; DeSauvage Nolting, H. J. J.; Mouljin, J. A. Fuel,

1988, 67, 334-340.

183 Cullis, C. F.; Norris, A. C. Carbon 1972, 10, 525-537.

184 Braslavsky, S.; Heicklin, J. Chem. Rev. 1977, 77, 473-511.

185 Klein, E.; Lukes, V.; Cibulkova, Z.; Polovkova, J. J. Mol. Struct. 2006, 758, 149-159.

103

186 Grela, M. A.; Amorebieta, V. T.; Colussi, A. J. J. Phys. Chem. 1985, 89, 38-41.

187 Organ, P. P.; Mackie, J. C. Chem. Soc. Faraday Trans. 1991, 87, 815-823.

188 Fulle, D.; Dib, A.; Kiefer, J. H.; Zhang, Q.; Yao, J.; Kern, R. D. J. Phys. Chem. A

1998, 102, 7480-7486.

189 Sendt, K.; Bacskay, G.B.; Mackie, J. C. J. Phys. Chem. 2000, 104, 1861-1875.

190 Lifshitz, A.; Tamburu, C.; Suslensky, A. J. Phys. Chem. 1989, 93, 5802-5808.

191 Bacskay, G. B.; Martoprawiro, M.; Mackie, J. C. Chem. Phys. Lett. 1998, 290, 391-

398.

192 Mackie, J. C.; Colket, M. B.; Nelson, P. F.; Esler, M. Int. J. Chem. Kinet. 1991, 23,

733-760.

193 Barckholtz, C.; Barckholtz, T. A.; Hadad, C. M. J. Am. Chem. Soc. 1999, 121, 491-

500.

194 Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6088-6094.

195 Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6324-6331.

196 Terentis, A.; Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10334-10339.

197 Ikeda, E.; Mackie, J. J. Anal. Appl. Pyrolysis 1995, 34, 47-63.

198 Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10339-10348.

199 Frerichs, H.; Schliephake, V.; Tappe, M.; Wagner, H. Z. Phys. Chem. 1990, 166.

200 Hayes, C.J.; Hadad, C.M. Combustion Pathways of the Alkyl-Substituted

Heteroaromatics: Bond Dissociation Enthalpies and Alkyl Group Fragmentations. J.

Phys. Chem. A, submitted.

104

201 (a) Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10339-10348. (b) Ebert, K. H.;

Ederer, H. J.; Schmidt, P. S., Baulch, D. L.; Cobos, C. J.; Cox, R. A.; Esser, C.; Frank, P.;

Just, T.; Kerr, J. A.; Pilling, M. J.; Troe, J.; Walker, R. W.; Warnatz, J. J. Phys. Chem.

Ref. Data 1992, 21, 411-429. (c) Wang, H.; Frenklach, M. J. Phys. Chem. 1994, 98,

11465-11489; (d) Hippler, H.; Troe, J. J. Phys. Chem. 1990, 94, 3803-3806. (e) Walker,

J.; Tsang, W. J. Phys. Chem. 1990, 94, 3324-3327. (f) Ellison, G. B.; Davico, G. E.;

Bierbaum, V. M.; DePuy, C. H. Int. J. Mass Spectrom. Ion Proc. 1996, 156, 109-131. (g)

Muralha, V.; dos Santos, R.; Simoes, J. J. Phys. Chem. A 2004, 108, 936-942.

202 Hayes, C. J.; Hadad, C. M. Combustion Pathways of the Alkyl-Substituted

Heteroaromatics: Peroxy Radical Pathways of the Alkylated Azabenzenes, J. Phys.

Chem. A, submitted.

203 Hayes, C. J.; Hadad, C. M. Combustion Pathways of the Alkyl-Substituted

Heteroaromatics: Peroxy Radical Pathways of the Alkylated Five-Membered

Heteroaromatics, J. Phys. Chem. A, submitted.

204 (a) Onishi, S.; Jo, S.; Shoda, K.; Jo, P. D.; Kato, S. SAE Paper #790501 (1979). (b)

Noguchi, M.; Tanaka, Y.; Tanaka, T.; Takeuchi, Y. SAE Paper #790840 (1979). (c)

Najt, P. M.; Foster, D.E. SAE Paper #830264 (1983). (d) Aceves, S. M.; Smith, J. R.;

Westbrook, C. K.; Iptz, W. J. J. Engin. Gas Turb. Power 1999, 121, 569-574. (e)

Westbrook, C. K. 28th Int’l Symp. Comb. 2000, 1563-1577. (f) Aceves, S. M.; Flowers,

D. L.; Martinez-Frias, J.; Smith, J. R.; Chen, J. Y. SAE Paper # 149614 (2002). (g)

Bunting, B.G. SAE Paper # 010872 (2006).

105

205 (a) Aceves, S.M.; Flowers, D.L.; Martinez-Frias, J.; Smith, J.R. SAE Paper #01-2077

(2001). (b) Wang, R.C.; Fang, T. SAE Paper #011519 (2006).

206 Thring, R. H. Homogeneous-Charge Compression-Ignition (HCCI) Engines.

Warrendale, PA: SAE, 1989

207 Carlsen, W. San Francisco Chronicle. 15 September 1997.

208 Marinov, N. M. Int. J. Chem. Kinet. 1999, 31, 183-220.

209 Energy Policy Act of 1992 (42 USC 10141 n). Online. Internet. Available

http://thomas.loc.gov/cgi-bin/query/z?c102:H.R.776.ENR. Accessed 15 February 2007.

210 DiPardo, J. “Outlook for Biomass Ethanol Production and Demand.” EIA Forecasts.

Online. Internet. Available http://www.eia.doe.gov/oiaf/analysispaper/biomass.html.

Accessed 6 July 2006.

211 Chase, R. “DuPont, BP join to make butanol.” USA TODAY. 30 June 2006.

212 “Butanol—Better than Ethanol as a Gasoline Substitute.” Online. Internet. Available

http://alfin2300.blogspot.com/2006/06/butanol-better -than-ethanol-as.html. Accessed 30

June 2006.

213 “Environmental Energy: Butanol Replaces Gasoline.” Online. Internet. Available

http://www.butanol.com/. Accessed 30 June 2006.

214 Ramey, D.; Yang, S.-T. Production of Butyric Acid and Butanol from Biomass. U.S.

Department of Energy: Morgantown, WV, 2004. Online. Internet. Available

http://www.osti.gov/bridge/servlets/purl/843183-rqlY2a/843183.PDF. Accessed 15 July

2006.

106

215 “Butanol.” Online. Internet. Available http://peswiki.com/index.php/Directory:Butanol. Accessed 15 July 2006.

216 Ramey, D. “Advances in Biofuels: Alternative Energy and New Markets for

Farmers.” Online. Internet. Available http://www.lightparty.com/Energy/Butanol.html.

Accessed 24 July 2006.

217 Marinov, N. M. Int. J. Chem. Kinet. 1999, 31, 183-220.

218 Cavalli, F.; Geiger, H.; Barnes, I.; Becker, K.H. Environ. Sci. Technol. 2002, 36,

1263-1270.

219 Chen, Y.L.; Chou, T.-C. Ind. Eng. Chem. Res. 1996, 35, 2172-2176.

220 McEnally, C.S.; Pfefferle, L.D. Proc. Combust. Inst. 2006, 30, 1363-1370.

221 Westbrook, C. K.; Pitz, W. J.; Curran, H. J. J. Phys. Chem. A 2006, 110, 6912-6922.

222 Sinha, A.; Thomson, M. J. Combust. Flame 2004, 136, 548-556.

223 Nag, P.; Song, K. H.; Litzinger, T. A.; Haworth, D. C. Int. J. Engine Res. 2001, 2,

163-175.

224 Bowman, F. M.; Seinfeld, J. H. Prog. Energy Combust. Sci. 1995, 21, 387-417.

225 (a) Schwab, A. W.; Bagby, M. O.; Freedman, B. Fuel 1987, 66, 1372-1378; (b)

Schwab, A. W.; Dykstra, G. J.; Selke, E.; Sorenson, S. C.; Pryde, E. H. JOACS 1988, 65,

1781-1786.

226 Graboski, M. S.; McCormick, R. L. Prog. Energy Combust. Sci. 1998, 24, 125-164.

227 Srivastava, A.; Prasad, R. Renew. Sust. Energ. Rev. 2000, 4, 111-133.

228 Lin, C.-Y.; Lin, H.-A.; Hung, L.-B. Fuel 2006, 85, 1743-1749.

229 Ma, F.; Hanna, M. A. Bioresour. Technol. 1999, 70, 1-15.

107

230 Kulkarni, M. G.; Dalai, A. K. Ind. Eng. Chem. Res. 2006, 45, 2901-2913.

231 Zhenyi, C.; Xing, J.; Shuyuan, L.; Li, L. Energy Sources 2004, 26, 849-856.

232 Gail, S.; Thomson, M. J.; Sarathy, S. M.; Syed, S.A.; Dagaut, P.; Dievart, P.;

Marchese, A. J.; Dryer, F. L. Proc. Combust. Inst. 2007, 31, in press.

233 Fisher, E.M.; Pitz, W.J.; Curran, H.J.; Westbrook, C.K. Proc. Combust. Inst. 2000, 28,

1579-1586.

234 Good, D. A.; Francisco, J. S. J. Phys. Chem. A 2000, 104, 1171-1185.

235 (a) Metcalfe, W. K.; Dooley, S.; Curran, H. J.; Simmie, J. M.; El-Nahas, A. M.;

Navarro, M. V.2007, 111, 4001-4014; (b) El-Nahas, A. M.; Navarro, M. V.; Simmie, J.

M.; Bozzelli, J. W.; Curran, H. J.; Dooley, S.; Metcalfe, W. J. Phys. Chem. A. 2007, 111,

3727-3739.

236 Glaude, P. A.; Pitz, W. J.; Thomson, M. J. Proc. Combust. Inst. 2005, 30, 1111-1118.

237 Schwartz, W. R.; McEnally, C. S.; Pfefferle, L. D. J. Phys. Chem. A 2006, 110, 6643-

6648.

238 Sarathy, S. M.; Gail, S.; Syed, S. A.; Thomson, M. J.; Dagaut, P. Proc. Combust. Inst.

2007, in press.

108

CHAPTER 2

COMBUSTION PATHWAYS OF THE ALKYLATED HETEROAROMATICS: BOND DISSOCIATION ENTHALPIES AND ALKYL-GROUP FRAGMENTATIONS

2.1 Introduction

Coal is a valuable fossil fuel, accounting for the majority of America’s electrical

power generation, and it is likely to become even more useful in the years ahead, as natural gas and oil reserves dwindle.1 The actual structure of a given coal specimen

varies greatly with geography;2 however, aromatic hydrocarbons and heteroaromatic

rings (Figure 2.1), both alkylated and non-alkylated, are recurring components, the

proportions of which differ between regions and samples.3

The practice of using the reactions of these individual rings to better understand

the overall chemistry of the complex structure of coal has been widespread over the past

few decades. In separate studies, Mackie et al. and Kiefer et al. have examined shock

tube studies of pyridine, the picolines (methyl-substituted pyridines), pyrazine, and

pyrimidine to better understand the thermal decompositions of coal units.4 These studies

have focused only on pyrolysis; thus, the understanding of the oxidative decompositions

of these heteroaromatics are still of interest. Moreover, Eisele established the presence of

109

pyridine and picoline ions in the troposphere, such that the low-temperature oxidation

pathways of these species are also relevant and could have implications for atmospheric

processes.5 These heteroaromatic rings have the potential to be oxidized at the nitrogen

6 atom and thus form NOx species, which can subsequently react to over-produce

tropospheric ozone and contribute to acid rain.7

H N O S O 5 2 2 2 2 4 3 3 3 N Pyrrole Furan Thiophene Oxazole N N N N 2 N 2 2

3 3 5 N N 4 4 4 Pyridine Pyridazine Pyrimidine Pyrazine Figure 2.1. Heteroaromatic rings of interest.

Our group has devoted much time to the study of these coal constituents.

Barckholtz et al. completed an exhaustive survey to select an appropriate computational method for study of these relatively large molecules; density functional theory (DFT)8 was shown to balance accurate results with computational economy to the greatest extent.9 This study also allowed the determination of a crucial conclusion, that the

110

monocyclic heteroaromatic rings do provide constructive models for comparison to their

polycyclic analogues, via comparisons of sp2 C-H bond dissociation enthalpy (BDE) values in polycyclic compounds with the corresponding BDE in the relevant monocyclic ring. The vast majority of these calculations showed that increasing the number of rings in the compound affected the BDEs by less than 1 kcal/mol, and thus the smaller rings provide worthy, computationally-prudent models for the larger systems. Moreover, it was shown that heteroatoms in the six-membered rings provided stabilization for the resulting radical, such that pyridine and other heteroaromatic rings had lower BDE values than did benzene alone.

The combustive pathways of these heteroaromatics have also been explored, at

multiple temperatures. Fadden et al. saw that there were considerable differences

between the pathways of the five-membered heteroaromatics’ peroxy radicals and the

azabenzenes’ peroxy radicals;10 the former could lose atomic oxygen at a relatively low

energetic cost, while for the six-membered rings, losing atomic oxygen was substantially

more unfavorable, to the extent that dissociation back to the aromatic radical and

molecular oxygen was preferable. For both sets of compounds, intramolecular

cyclizations competed with oxygen loss, and some of these cyclizations did lead to

nitroso compounds, which have implications both for NOx formation and as

carcinogens.11

Computational and experimental methods clearly benefit from a symbiotic

relationship in combustion studies: calculations can propose pathways for empirically-

observed intermediates, and experiments can verify the accuracy of the calculated

111

energies and rate coefficients. Combustion models have been proposed for several

classes of molecules, including alkyl hydrocarbons and aromatic species.

To this point, little has been published on the subject of alkylated heteroaromatic

rings, though these species demonstrate comparable promise for elucidating coal

chemistry. Notably, Mackie and coworkers completed both experimental12,13 and

theoretical14 studies of the pyrolytic decomposition of 2-picoline (2-methylpyridine).

However, more than a decade later, this remains the most exhaustive study of an

alkylated heteroaromatic ring’s combustion pathway, although the authors hypothesized

that these actually provided a better model for fuel-bound nitrogen (FBN) than did

unsubstituted heteroaromatic rings, which have been more common targets.- The kinetics

of the methyl-substituted azabenzenes’ initial reactivities have been explored more

frequently: Frerichs et al. examined the reaction of the picolines with atomic oxygen,15 while Yeung and Elrod completed a study of hydroxyl-radical’s reactions with pyridine and its methyl and ethyl substituted derivatives.16 Both groups noted similarities

between the alkylated azabenzenes and the reactivity of toluene.

Unsurprisingly, with respect to toluene and the other alkylated aromatic

hydrocarbons, more references are available. Both experimental and theoretical studies

have been completed on the combustion processes of toluene. Most recently, Pitz et al.

published a comprehensive mechanistic study using shock-tube techniques;17 earlier studies had dealt specifically with improving predictions of some of the key intermediates therein.18 Computationally, substituent effects are a popular area of study as well as the

overall mechanism: the hydrogen-atom loss and subsequent atmospherically relevant

112

reactions of toluene have been studied via semi-empirical methods19 and the B3LYP

functional.20 Nam et al. completed a DFT study of toluene and its para and meta-

substituted derivatives, observing how substituents on the methyl group itself further

impacts the bond dissociation enthalpies of these species.21 Ethylbenzene has been

subjected to several studies; both its pyrolysis22 and its reactions with atomic oxygen,23 fluorine radical,24 hydrogen atom,25 hydroperoxyl radical,26 and benzyl radical27 have

been studied. These hydrocarbon compounds can provide references for energies and

likely pathways for the corresponding alkylated heteroaromatic rings.

Of greatest interest to this particular work are the bond dissociation enthalpies

(BDEs) of toluene and ethylbenzene. Besides their use as reference compounds for the

substituted heteroaromatics, they are also of interest as intermediates in the HACA (H-

28 abstraction-C2H2-addition) pathway to soot formation. A necessary first step in any

combustion pathway is generation of a radical species that can be subsequently oxidized,

and this homolytic cleavage is represented quantitatively via BDEs. The BDEs of

toluene’s methyl group and ethylbenzene’s benzyl position have been experimentally

determined by several methods, including flowing-afterglow mass spectrometry, shock

tubes, and photoacoustic calorimetry; these were recently compiled by Muralha et al.29

The BDE values calculated for toluene varied from 88.0 kcal/mol30 to 90.3 kcal/mol,31 with the majority of results falling in the 89.5-90 kcal/mol range. With ethylbenzene, the resulting radical saw stabilization from the additional alkyl group, so its BDE was lower, ranging from 85.4 kcal/mol to 86.9 kcal/mol.,32 ,33

113

In general, bond dissociation enthalpies are useful quantities in terms of

predicting relative reactivities for different compounds, reflecting the enthalpy change for

homolytic cleavage of a given bond. The more stable the resulting radical, the more

likely a bond is to break. General explanations for radical stability include alkyl

substitution and resonance effects; Gronert has recently suggested an alternative

explanation, that the potential for release of 1,3-repulsive energy (strain) has the greatest effect on these quantities: i.e., a tertiary radical is more stable than a primary radical not because of the larger number of alkyl substituents in the radical, but because of the greater magnitude of the geminal interactions in the parent hydrocarbon that are relieved upon C−H bond cleavage.34

The BDEs of the alkylated heteroaromatics are expected to reflect the resonance

stability of the formed radicals via conjugation to the aromatic system and the

stabilizing/destabilizing role of the relevant heteroatom(s). Hydrogen-atom loss seems to

be the most likely initiation step in the combustion of the alkylated heteroaromatic rings,

although other pathways, such as loss of the alkyl group, are also plausible (Figure 2.2).35

114

X X CH3 X CH2

+CH 3 + H

X X CHCH X CH2CH3 3

+CH2CH3 + H

Figure 2.2. Generic representation of possible radical-generation pathways for the alkylated heteroaromatic rings.

These cleavage pathways will be chain-propagating for the combustion process and may dominate at higher temperatures due to the favorable entropic term.

Just as alkylated aromatic rings are widespread in petroleum compounds, alkylated heteroaromatic rings are prevalent in coal compounds, and their combustion pathways are thus noteworthy. Thus, the hydrogen-atom loss reactions and alkyl-radical loss reactions for the methyl and ethyl-substituted five- and six-membered heteroaromatic rings will be explored via DFT methods, with the goal of further elucidating the chemistry of the larger heterocyclic systems found in coal. Additionally, the DFT results will be calibrated against CBS-QB3 data, with the intent of validating the B3LYP method as an economical method with which to further explore the oxidative pathways of the consequent radicals.

115

2.2 Computational Methods

All geometry optimizations, vibrational frequency calculations, and single-point

energy calculations were completed with Gaussian 9836 and 0337 at the Ohio

Supercomputer Center. The B3LYP/6-31G* hybrid density functional theory (DFT)38 level was used for all geometry optimizations and vibrational frequency calculations.

Single-point energies were calculated at the B3LYP/6-311+G** level, using six Cartesian d functions with the scf=tight option; it has been shown that there is a small basis set effect with more polar systems,9,39 so this larger basis set was selected to compensate for

that effect. CBS-QB3 calculations40 were also performed to validate the DFT method for

these particular systems.

Vibrational frequencies were calculated for each stationary point to characterize

these structures as minima or transition states. The unscaled vibrational frequencies were

used to calculate the thermodynamic corrections to the enthalpy and free energy. Once

obtained, zero-point vibrational energy corrections were scaled by a factor of 0.9806.41

The overall enthalpy at each temperature was determined from the single-point energy, the thermal correction to the enthalpy, and the scaled zero-point energy, while the overall free energy at each temperature also included the entropic correction to the free energy.

Transition states were connected to reactants and products by either using intrinsic reaction coordinate (IRC)42 searches or displacing the relevant geometries by

±10% along the reaction coordinate characterized by the imaginary vibrational frequency,

116

then carefully optimizing (opt = calcfc) these resulting geometries to their corresponding

minima.

For the doublet radical species and many transition states, spin contamination

() was negligible (no greater than 0.80). The DFT methods employed in this work have demonstrated an ability to minimize spin contamination for the radicals of interest,

keeping the computed values near the expected value of 0.75. Spin densities were

obtained for the hydrogen-loss radicals using the natural population analysis (NPA)

method.43

Because flame processes are some of the reactions of interest for these

compounds, it is necessary to see how these species react at higher temperatures. The

enthalpies of both the parent compounds and the resultant radicals were calculated from

298 to 2000 K via the temperature-dependent term:44

=Δ trans + rot + vib )()()()( + RTTHTHTHTH

3 3 i / kThv −1 )( ++=Δ ∑ i evNhRTRTTH )1( +− RT 2 2 i

The summation takes place over all 3N-6 normal modes. For each compound, the scaled

zero-point energy and enthalpy contribution were computed at various temperatures via the given equations and added to the B3LYP total energy, to obtain the bond dissociation enthalpies as a function of temperature. A similar approach was used with calculation of the entropic corrections, to determine the free energies of reaction over the same temperature range.45

To reduce potential errors in the thermodynamic values arising from treating all

vibrations according to the harmonic-oscillator rigid-rotor approximation, low-energy

117

torsions in the alkylated species were also treated as hindered rotors,46 and the resulting

energetics were compared to those from the original calculations.

2.3 Results

Several alkylated heteroaromatic rings were examined: pyrrole, furan, thiophene,

oxazole, pyridine, pyrimidine, pyridazine, and pyrazine (Figure 2.1). Bond dissociation

enthalpies (BDEs) and energies were compiled for the benzyl C−H BDE in both the methyl and ethyl derivatives of these compounds, as were the corresponding spin densities in the derived radicals. Enthalpies and free energies of reaction for loss of the attached alkyl groups were also documented. BDE values were then analyzed both as a function of excess spin density on the major radical center of interest and of temperature.

Finally, the harmonic oscillator and rigid rotor approximations were compared.

2.4 Discussion

Several trends can be observed from the enthalpic and free energy data; for

simplicity’s sake, these trends will be discussed in terms of their relative bond

dissociation enthalpies. Moreover, the nomenclature used will indicate that all radicals

are located on the benzylic position of the compound; that is, 2-methylpyridinyl radical

• will be used as an abbreviation for 2-methylpyridin-2’-yl (c-C5H4N-CH2 ) radical, throughout the remainder of the paper.

With respect to the different methods, the hydrogen-loss BDE values were

calculated and tabulated for both the CBS-QB3 and B3LYP calculations. The DFT

118

enthalpies and energies of reaction were consistently 3−4 kcal/mol lower than the CBS results (which were themselves comparable to experimental results, where available); thus, the qualitative trends were duplicated between methods. The majority of the comparative discussion will refer back to the B3LYP geometries and energies.

2.4.1 Toluene as reference compound

The BDE and geometry of toluene were both explored, along with the substituted heteroaromatic rings. Since more information is available for toluene, it can serve as a standard for evaluating the accuracy of the computational approach, which can then be applied to the heteroaromatic rings as well.

As shown in the Supporting Information, the B3LYP/6-31G* method matched experimental values for toluene’s geometry very well: both the C-C and C-H bond lengths are reproduced when compared to the spectroscopic work of Amir-Ebrahimi et al.,47 and the rotational constants similarly match up. The vibrational frequencies, also reported in the Supporting Information, saw a comparable correspondence.

Comparable information on the benzyl radical was not readily available; however, it can be inferred from the bond dissociation enthalpy of toluene, as both parent and radical must necessarily be geometrically correct to give an accurate BDE value. The

CBS-QB3 calculations predict a BDE of 90.6 kcal/mol, matching well with the experimental data (a range of 88.0 – 90.3 kcal/mol). The B3LYP/6-311+G**//B3LYP/6-

31G* energies ultimately provide a calculated BDE of 86.7 kcal/mol; the 4 kcal/mol difference from the CBS-QB3 calculation is consistent throughout the heteroaromatic

119

rings studied, such that qualitative trends are reproduced. The slight underestimate is

common to the DFT approach and is seen in other work in this area.9

2.4.2 Non-alkylated heteroaromatics

2.4.2.1 Azabenzenes: Just as this work uses toluene and ethylbenzene as benchmarks for

the alkylated heteroaromatic rings, previous work on the unsubstituted heteroaromatic rings has shown several trends in their BDEs relative to that of benzene. With respect to the azabenzenes, it was seen that the presence of one or more nitrogen atoms within the aromatic ring directly affected those compounds’ BDEs. In particular, the BDE at the C-

H bond adjacent to nitrogen was consistently seen to be around 4−5 kcal/mol lower than benzene’s C-H BDE, a phenomenon attributed to nitrogen’s ability to lend greater resonance stabilization to the heterocyclic radical. At the non-nitrogen-adjacent positions, the C-H BDE was comparable to that of benzene itself.

2.4.2.2 Five-membered rings: Prior work on the unsubstituted five-membered heteroaromatic rings has shown that their combustion pathways are most likely to begin with decomposition or isomerization of the ring itself, rather than loss of hydrogen atom.

This is due to the geometric perturbation introduced by forming a radical centered on the

aromatic ring, which is reflected in the higher C−H BDE values compared to benzene.

Furthermore, unlike the azabenzenes, there is no distinct preference of H-abstraction from position 2 over position 3; again, due to geometric factors, the heteroatoms are unable to stabilize the radical to the degree seen in the six-membered rings.

120

Shifting focus to the alkylated heteroaromatic rings, it seems probable that the

chemistry here could differ substantially, since these H-loss reactions will not create a

radical center within the aromatic ring itself, but rather on the attached alkyl group. This

proves to be the case, as these trends differ from those of the unsubstituted compounds.

2.4.3 Methyl-substituted heteroaromatics: Hydrogen-atom loss

2.4.3.1 Azabenzenes: The methyl-substituted azabenzenes vary in their BDEs from 86.3

to 92.6 kcal/mol, with an average value of 87.2 kcal/mol, higher than toluene’s value of

86.7 kcal/mol (Table 2.1). The experimental work of Doughty and Mackie on 2- methylpyridine14 did provide a tangential reference point here; their proposed C−H BDE

was 96 kcal/mol, compared to the calculated value of 88.2 kcal/mol via B3LYP (92.0

kcal/mol via CBS-QB3). Calculations were seen to underestimate the BDE by a slightly

greater margin than in toluene’s case. The authors noted that the heat of formation for

their 2-picolinyl radical was an estimate, which would extend to the extrapolated BDE of interest and could explain this discrepancy. Losing hydrogen from the attached methyl group rather than from the ring itself means that the resultant radical is not directly on the ring, and not as directly stabilized by nitrogen’s additional resonance effects. Moreover,

the trend of the ortho-substituted compound’s BDE being lower than the other substituted

compounds is not duplicated here: in fact, 2-methylpyridine and 3-methylpyridazine both

have substantially higher BDEs than their counterparts, and 2-methylpyrimidine’s BDE

value is near the highest value for the pyrimidine derivatives.

121

X CH3 X CH2

+ H

ΔH298 ΔG298 (α - β) Methyl B3LYP CBS Experimental B3LYP CBS BDE Toluene 1 86.7 90.6 88.0-90.3b 79.4 83.8 0.72 Pyrrole 2 83.1 86.1 75.3 78.9 0.63 3 86.8 90.1 78.9 82.4 0.74 Furan 2 83.1 86.3 75.3 78.6 0.60 3 87.4 90.5 79.5 82.7 0.73 Thiophene 2 85.2 86.5 84.2 79.1 0.60 3 87.0 89.9 88.1 82.2 0.71 Oxazole 2 86.4 89.9 78.7 82.3 0.64 4 88.1 91.1 80.2 83.3 0.63 5 84.5 87.9 76.6 80.1 0.72 Pyridine 2 88.2 92.0 96.0c 80.5 84.7 0.73 3 87.0 91.0 79.6 83.7 0.72 4 87.9 91.6 80.6 84.6 0.74 Pyridazine 3 88.9 93.3 81.3 85.8 0.76 4 87.6 91.7 80.5 84.6 0.75 Pyrimidine 2 89.5 93.1 82.4 86.7 0.73 4 89.3 92.7 81.9 85.5 0.75 5 87.5 91.4 80.7 84.7 0.72 Pyrazine 2 88.0 92.3 80.5 85.0 0.72 a Enthalpies and energies in kcal/mol, obtained via B3LYP/6-311+G**// B3LYP/6-31G* (designated as B3LYP) and CBS-QB3 (designated as CBS) methods. See Figure 2.1 for structures and numbering. b References 24-26 c Reference 13

Table 2.1: Thermodynamic information and spin density information for hydrogen-loss (C-H homolytic bond cleavage) reactions of methyl-substituted heteroaromatic rings. a

122

Both geometries and spin densities were examined to discern the origin of these effects. The geometry changes involved are minimal, as shown (Table 2.2). Virtually the same changes occur for toluene losing a hydrogen atom from the methyl group as for 2- methylpyridine.

1 1 6 6 CH3 CH2 N CH3 N CH2 5 5 6 6 1 1 2 2 + H + H 4 2 4 2 5 3 5 3 3 3 4 4

Bond Lengths Parent Radical Change Parent Radical Change C6-C1 1.401 1.427 0.026 N1-C2 1.345 1.372 0.027 C1-C2 1.401 1.427 0.026 C2-C3 1.401 1.425 0.024 C2-C3 1.396 1.389 -0.007 C3-C4 1.394 1.389 -0.005 C3-C4 1.396 1.403 0.007 C4-C5 1.393 1.399 0.006 C4-C5 1.396 1.403 0.007 C5-C6 1.396 1.405 0.009 C5-C6 1.396 1.389 -0.007 C6-N1 1.337 1.326 -0.011 C1-Cmeth 1.512 1.407 -0.105 C2-Cmeth 1.509 1.408 -0.101 Bond Angles Parent Radical Change Parent Radical Change C6-C1-Cmeth 121.4 121.4 0 N-C2-Cmeth 116.3 117.2 0.9 C2-C1-Cmeth 120.1 121.4 1.3 C3-C2-Cmeth 121.7 121.6 -0.1 C1-Cmeth-H 110.5 121.2 10.7 H-Cmeth-C2 111.8 C1-Cmeth-H 110.1 121.2 11.1 H-Cmeth-C2 110.4 119.6 9.2 C1-Cmeth-H 112.3 H-Cmeth-C2 110.4 121.3 11.1

Table 2.2. Comparison of changes in bond lengths (Angstroms) and bond angles (degrees) during the hydrogen-atom loss reactions of toluene and its nitrogen analogue, 2-methylpyridine.

123

Spin density (population) appears to be the most logical candidate for rationalizing these

trends. From the relative electron distribution (α-β) between benzyl radical and 2-

methylpyridinyl radical (Figure 2.3), it can be seen that 2-methylpyridinyl radical

localizes its electron density on the CH2 carbon to a greater extent than benzyl radical;

correspondingly, benzyl radical sees more electron density delocalized through the ring

and a resultant stabilization. In particular, the excess spin density is localized on the

positions ortho and para to the attached methyl group, and the nitrogen of 2-

methylpyridinyl radical bears less spin density than does the corresponding carbon in

benzyl radical. Similar trends were seen for the other methyl-substituted heteroaromatics

(and will be discussed subsequently), although this is the most direct comparison

possible. Interestingly, the ethyl-substituted heteroaromatic rings’ trends are reversed

relative to ethylbenzene.

0.178 (0.215) 0.730 (0.718) 0.177 0.672 (0.196) (0.684)

0.214 0.236 (0.215) 0.229 0.198 (0.242) (0.224) (0.208)

Figure 2.3. Comparison of areas of increased spin densities (α-β) for 2-methylpyridinyl radical and 2-ethylpyridinyl radical. The top number refers to the α-β value for that particular position; the bottom number, in parentheses, refers to the corresponding α-β value for the hydrocarbon equivalent (toluene and ethylbenzene, respectively).

124

2.4.3.2 Five-membered rings: In another marked difference from the unsubstituted

heteroaromatic rings, the five-membered, methyl-substituted heteroaromatic rings

consistently demonstrate BDEs lower than any of their six-membered counterparts,

including toluene. Because the radical center is removed from the ring, the geometric

perturbations that affected the unsubstituted compounds are no longer a major concern.

Moreover, loss of hydrogen atom from the methyl group forms an allylic radical, which

can now maintain the planarity necessary for delocalization.

The 2-methyl-substituted, five-membered heteroaromatic rings have lower BDEs

than their 3-methyl-substituted counterparts. Presumably, the trend seen in the

unsubstituted azabenzenes helps to rationalize the results seen here: the compounds in

which the methyl group is ortho to the heteroatom are more able to exploit the additional

resonance stability afforded by nitrogen, oxygen, and sulfur relative to carbon.

Heteroatom identity does not have a large effect on energetics for the five-

membered rings, the only set of compounds for which the heteroatom was varied. The 2-

methyl-substituted heteroaromatic rings, in particular, have virtually identical BDEs. In the methyl derivatives of oxazole, the heteroaromatic containing both nitrogen and oxygen, it appears that the nitrogen has a more stabilizing effect than oxygen on the resulting radical, as the 4-methyl derivative (ortho to nitrogen) has a lower BDE than the

2-methyl derivative (ortho to oxygen).

125

2.4.4 Ethyl-substituted heteroaromatics: Hydrogen-atom loss

Overall, the ethyl-substituted heteroaromatic rings have considerably lower BDEs than their methyl counterparts; the additional alkyl group serves to stabilize the radical derived from hydrogen-atom loss, regardless of ring size, by 4 to 5 kcal/mol (Table 2.3).

ΔH298 ΔG298 (α - β) Ethyl B3LYP CBS Experimental B3LYP CBS BDE Ethylbenzene 1 83.9 88.1 85.4-86.9b 75.2 79.7 0.68 Pyrrole 2 80.8 84.6 72.2 75.9 0.61 3 83.8 87.7 75.3 78.9 0.71 Furan 2 80.1 83.9 71.6 75.5 0.57 3 84.2 87.9 75.4 79.1 0.69 Thiophene 2 79.8 83.8 71.1 75.2 0.55 3 83.5 87.1 74.4 78.1 0.67 Oxazole 2 82.4 86.5 74.1 78.4 0.61 4 84.5 88.3 76.0 79.9 0.68 5 81.5 85.6 73.0 77.1 0.60 Pyridine 2 84.2 88.1 75.2 80.0 0.67 3 82.9 85.6 74.8 77.2 0.69 4 83.7 87.2 74.5 78.7 0.69 Pyridazine 3 84.7 90.1 77.6 83.1 0.68 4 83.7 88.7 75.1 79.7 0.68 Pyrimidine 2 84.2 89.0 76.1 80.9 0.69 4 83.9 88.5 75.7 80.4 0.70 5 84.3 89.1 74.9 80.5 0.66 Pyrazine 2 83.8 90.2 75.1 81.4 0.68 a Enthalpies and energies in kcal/mol, obtained at the B3LYP/6-311+G**//B3LYP/6-31G* (designated as B3LYP) and CBS-QB3 (designated as CBS) levels. See Figure 2.1 for structures and numbering. b References 24, 27, and 28

Table 2.3: Thermodynamic and spin density information for reactions of ethyl- a substituted heteroaromatic rings, for hydrogen loss from the CH2 group.

126

2.4.4.1 Azabenzenes: The BDE values for the ethyl-substituted azabenzenes actually

encompass the value calculated for ethylbenzene (83.9 kcal/mol), as they range from 81.8

to 84.7 kcal/mol. The narrower range and lack of any trend relative to ethylbenzene

suggest that the additional methyl group present stabilizes all of the relevant radicals to a comparable extent.

2.4.4.2 Five-membered rings: The same trend seen with the methyl-substituted heteroaromatics is repeated here: the 2-substituted compounds have substantially lower

BDEs, again due to their increased proximity to the stabilizing heteroatom. Moreover, these compounds demonstrate more of a heteroatom effect, although it is a slight trend: the larger the heteroatom, the lower the BDE. Presumably, polarizability effects in stabilizing the radical might become more visible as the number of electrons in the molecule increase.

2.4.5 Methyl-substituted heteroaromatic rings: Methyl loss

Scission of the Cring-Calkyl bond for all of these compounds leads to an in-plane,

sp2 aromatic radical, localized in the σ system of the aromatic ring. It was expected that

the thermodynamics of these reactions would compare well to those for H-atom loss from

the unsubstituted heteroaromatic rings, and this was confirmed to be the case (Table 2.4).

Moreover, the thermochemistry for the loss of these alkyl groups was considerably more

endothermic and endoergic, in all cases, than hydrogen-atom loss.

127

Methyl ΔH298 ΔG298 Ethyl ΔH298 ΔG298 Pyrrole 2 106.1 89.3 2 101.1 87.1 3 104.3 87.7 3 99.1 85.6 Furan 2 108.8 92.1 2 103.9 89.9 3 105.7 89.1 3 100.6 86.8 Thiophene 2 103.9 87.3 2 99.1 85.3 3 100.9 84.2 3 95.6 81.6 Oxazole 2 109.8 93.3 2 104.9 91.1 4 106.9 90.3 4 102.1 88.3 5 110.8 94.1 5 105.9 92.1 Pyridine 2 94.1 77.9 2 89.3 75.5 3 98.3 82.1 3 93.4 79.4 4 97.6 81.7 4 92.7 78.9 Pyridazine 3 96.2 79.7 3 91.4 77.7 4 97.1 81.2 4 92.2 78.6 Pyrimidine 2 97.4 81.5 2 92.5 78.8 4 90.3 74.1 4 89.7 75.9 5 99.7 84.1 5 94.9 81.1 Pyrazine 2 94.9 78.5 2 87.1 72.3 a Enthalpies and energies in kcal/mol, obtained at the B3LYP/6-311+G**//B3LYP/6-31G* level of theory.

Table 2.4: Thermodynamic information for alkyl-group loss reactions of methyl- and ethyl-substituted heteroaromatics.a

2.4.5.1 Azabenzenes: Loss of the methyl group was seen to be around 10 kcal/mol more endothermic than loss of hydrogen, for any given position in these compounds.

Moreover, when the methyl group in question was ortho to nitrogen (as with 2- methylpyridine compared to the other pyridine derivatives, and 3-methylpyridazine

128

compared to 4-methylpyridazine), the corresponding C-C BDE was even lower, due to

the stabilizing effects of the adjacent nitrogen. With the pyrimidine derivatives, the trend

is still evident; 4-methylpyrimidine, with its methyl group between two nitrogens, has a

substantially lower BDE than any other in this group.

2.4.5.2 Five-membered rings: The trends seen for hydrogen-loss reactions are non-

existent here. As seen in earlier work by our group9 and others,48 any stabilizing effects of the nearby heteroatom are negated by the geometric strain introduced by forming a ring-centered radical: the BDE at the 2-position is generally 2 to 3 kcal/mol higher than at

the 3-position. No definite heteroatom trends are evident, and the oxazole derivatives

demonstrate that the presence of two heteroatoms does not lead to greater reactivity at any one position.

2.4.6 Ethyl-substituted heteroaromatic rings: Ethyl loss

The ethyl-substituted compounds (Table 2.4) have the same trends as the methyl

derivatives and so will not be discussed further, save to note that the BDE values are

again around five kcal/mol lower than those of the methyl compounds, probably due to

the increased stability of the primary ethyl radical over the methyl radical.

2.4.7 General trends

2.4.7.1 Spin density: The spin densities were calculated via Natural Population Analyses

at the B3LYP/6-311+G**//B3LYP/6-31G* level. Excess spin density (α – β) was

determined for each of the incipient CH2 radical centers, and BDE was plotted as a

129

function of this quantity (Figure 2.4). A comparable graph of BDE as a function of excess spin density was completed for the ethyl-substituted compounds (Figure 2.5).

Figure 2.4. Variation of bond dissociation enthalpy (kcal/mol) with excess spin density at the incipient CH2 radical center, for methyl-substituted heteroaromatic rings. Data for the five-membered heteroaromatic rings are denoted by closed circles; data for the six- membered heteroaromatic rings are denoted by open squares.

130

Figure 2.5. Variation of bond dissociation enthalpy (kcal/mol) with excess spin density at the incipient C-H radical center, for ethyl-substituted heteroaromatics. Data for the five-membered heteroaromatic rings are denoted by closed circles; data for the six- membered heteroaromatic rings are denoted by open squares.

The spin density generated at the incipient radical correlated well with the BDE regardless of ring size or substitution, suggesting that the more localized a resulting radical will be, the larger the BDE value (or the more that electron delocalization will stabilize a given radical, the lower the BDE for its corresponding parent compound will be). The ethyl derivatives saw a slightly lower degree of correlation (R2 = 0.79) than did

the methyl derivatives (R2 = 0.84).

The correlation between the methyl and ethyl derivatives’ respective bond

dissociation enthalpies, for all the heteroaromatic rings considered together, was minimal.

131

With the azabenzenes, the ethyl derivatives were far more uniformly stabilized than their

methyl counterparts; their correlation was thus poor. When only the five-membered rings

were considered (Figure 2.6), a substantially higher correlation was seen (R2 = 0.78).

Figure 2.6. Correlation between bond dissociation enthalpies for methyl and ethyl- substituted derivatives of the five-membered heteroaromatics.

2.4.7.2 Geometry: Several geometric changes occurred within these compounds, which can be understood by examining a representative reaction (Table 2.5) with 2-methylfuran.

132

1 1 5 O 5 O 2 2 CH3 CH2 + H 4 4 3 3

Compound Radical Change in O-C2 1.371 1.390 0.019 C2-Cmeth 1.479 1.377 -0.102 C2-C3 1.364 1.411 0.047 C3-C4 1.436 1.411 0.025 C4-C5 1.359 1.373 -0.14 C5-O 1.366 1.361 -0.005 Cmeth-H 1.097 1.083 -0.14

Compound Radical Change in bond angle O-C2-Cmeth 116.64 118.93 2.29 C3-C2-Cmeth 133.82 133.17 -0.65 H1-Cmeth-C2 111.57 H2-Cmeth-C2 111.57 120.93 9.36 H3-Cmeth-C2 109.87 120.06 10.19 H1-Cmeth-H2 108.17 H2-Cmeth-H3 107.36 119.01 11.65 H3-Cmeth-H1 108.16

Table 2.5. Changes in bond lengths (Angstroms) and bond angles (degrees) for hydrogen-loss reaction of representative heteroaromatic ring, 2-methylfuran.

In terms of bond lengths, the main change occurred in the bond between the furan ring

and alkyl chain, which shorted substantially as the bond took on more π character in the corresponding π radical. Bonds within the ring lengthened to some extent. With respect to bond angles, the O-C2-Cmeth angle increased slightly, and the H-Cmeth-C2 angles

133

increased substantially, as hybridization at the radical center changed from sp3 to sp2.

The analogous geometries of interest for the other heteroaromatics are summarized in the

Supporting Information. What is most significant is that, for these alkylated aromatics, the ring itself was relatively unperturbed, relative to previous studies in which the new radical was located directly on the heteroaromatic of interest.9

2.4.7.3 Temperature effects: Calculations were completed to examine the changes in

hydrogen-loss BDE at combustion temperatures for these compounds. As shown

(Figures 2.7 and 2.8), minor quantitative changes (increases of roughly 4 kcal/mol) in the

hydrogen-loss BDEs of the heterocyclic compounds occur over this temperature range;

the most dramatic variation, seen with 2-methylpyrrole and 3-methylpyrrole, is still less

than 10 kcal/mol. Thus, the 298 K thermodynamic data can also be useful in terms of predicting higher-temperature chemistry. It is notable that although any quantitative changes are relatively small, the qualitative trends are not completely consistent. Most of the heteroaromatics see a slight increase in their BDE value proportional to temperature, but a few actually exhibit slightly lower BDEs at higher temperatures (most evident in the case of 2-ethylpyrazine). These variations are minor; overall, chemistry at 298 K provides a good reference for the reactivity of the alkylated compounds.

134

(a) (b) (b) (a)

Figure 2.7. Variation of bond dissociation enthalpy (kcal/mol) with temperature (K) for hydrogen-atom loss in the (a) five-membered methyl-substituted heteroaromatics and (b) six-membered methyl-substituted heteroaromatics. The atypical trends of 2- methylpyrrole and 3-methylpyrrole (x) are denoted by an arrow.

135

(a) (b)

Figure 2.8. Variation of bond dissociation enthalpy (kcal/mol) with temperature (K) for hydrogen-atom loss in the (a) five-membered ethyl-substituted heteroaromatics and (b) six-membered ethyl-substituted heteroaromatics. The atypical trend of 2-methylpyrazine (-) is denoted by an arrow.

136

The free energies of reaction for hydrogen-atom loss were also explored as a function of temperature (Figures 2.9 and 2.10); these quantities experienced a large temperature effect, due to the entropic term for these dissociative processes. At high temperatures

(2000 K), the free energies of reaction are roughly 60 kcal/mol more favorable than at

298 K, making these radicals ready participants in high-temperature combustion processes.

(a) (b)

Figure 2.9. Variation of free energy of reaction (kcal/mol) with temperature (K) for hydrogen-atom loss in the (a) five-membered methyl-substituted heteroaromatic rings and (b) six-membered methyl-substituted heteroaromatic rings.

137

(a) (b)

Figure 2.10. Variation of free energy of reaction (kcal/mol) with temperature (K) for hydrogen-atom loss in the (a) five-membered ethyl-substituted heteroaromatic rings and (b) six-membered ethyl-substituted heteroaromatic rings.

Moreover, comparative temperature profiles were generated to explore the relative energetics of hydrogen-atom loss and alkyl-group loss (Figures 2.11 and 2.12).

Representative cases are shown; a more complete set of information is included in the

Supporting Information. Both pathways see large entropic contributions at higher temperatures. In all cases, hydrogen-atom loss is initially the favored pathway, but at

2000 K, alkyl-group loss is favored slightly. At no temperature is one pathway overwhelmingly dominant over the other, and the energetics actually overlap at several points, so that the further combustion pathways of both sets of radicals will likely contribute in comparable capacity.

138

(a) (b)

(c) (d)

Figure 2.11. Comparison of hydrogen-atom loss and methyl-group loss free energies with increasing temperature. Representative pathways shown for (a) 2-methylpyrrole, (b) 3-methylpyrrole, (c) 3-methylpyridazine, and (d) 4-methylpyridazine. Free energies for hydrogen-atom loss are denoted with open squares; free energies for methyl loss are denoted with solid triangles.

139

(a) (b)

(c) (d)

Figure 2.12. Comparison of hydrogen-atom loss and ethyl-group loss free energies with increasing temperature. Representative pathways shown for (a) 2-ethylpyrrole, (b) 3- ethylpyrrole, (c) 3-ethylpyridazine, and (d) 4-ethylpyridazine. Free energies for hydrogen-atom loss are denoted with open squares; free energies for ethyl loss are denoted with solid triangles.

2.4.7.4 Approximations: Another set of calculations was run to determine what, if any,

error was introduced via using the harmonic oscillator approximation in finding the

enthalpic and entropic thermal corrections. As described previously, the lowest-energy

vibrations (i.e. the motions of the alkyl chain) were identified and treated as rotations, using a hindered rotor approximation rather than the harmonic oscillator. The enthalpies are essentially consistent between approximations (Table 2.6), while the free energies reflect a small change, generally less than 1 kcal/mol. Considering that this variation in

140

approximation would have the most impact on the entropy of these species, it seems logical that the free energies would see more of an effect than the enthalpies. However, neither quantity was greatly affected.

ΔH ΔH ΔG ΔG ΔH ΔH ΔG ΔG Methyl 298 298 298 298 Ethyl 298 298 298 298 (vib) (rotor) (vib) (rotor) (vib) (rotor) (vib) (rotor) Pyrrole 2 83.1 83.1 75.3 74.5 2 80.8 80.5 72.2 72.7 3 86.8 86.8 78.9 79.6 3 83.8 83.7 75.3 75.5 Furan 2 83.1 83.1 75.3 76.0 2 80.1 79.9 71.6 72.8 3 87.4 87.4 79.5 80.3 3 84.2 84.1 75.4 75.8 Thiophene 2 85.2 85.2 77.4 78.0 2 79.8 79.5 71.1 71.7 3 87.0 87.0 79.1 79.8 3 83.5 83.5 74.4 73.8 Oxazole 2 86.4 86.4 78.7 79.4 2 82.4 82.3 74.1 74.8 4 88.1 88.0 80.2 80.8 4 84.5 84.5 76.0 76.2 5 84.5 84.5 76.6 77.4 5 81.5 81.5 73.0 72.3 Pyridine 2 88.1 88.3 80.5 81.0 2 84.2 84.3 75.2 78.0 3 87.0 87.3 79.6 80.1 3 82.9 82.9 74.8 74.7 4 87.9 88.1 80.6 80.9 4 83.7 83.9 74.5 75.8 Pyridazine 3 88.9 89.0 81.3 81.9 3 84.7 85.4 77.6 75.9 4 87.6 87.9 80.6 80.7 4 83.7 83.6 75.1 75.2 Pyrimidine 2 89.5 89.7 82.4 82.6 2 84.2 84.5 76.1 76.4 4 89.3 89.5 81.9 82.4 4 83.9 84.0 75.7 76.2 5 87.5 87.8 80.7 80.7 5 84.3 84.3 74.9 76.0 Pyrazine 2 88.0 88.2 80.5 81.1 2 83.8 83.7 75.1 75.7 a Enthalpies and energies in kcal/mol, obtained at the B3LYP/6-311+G**//B3LYP/6- 31G* level of theory.

Table 2.6: Comparison of approximations (harmonic oscillator to hindered rotor), as applied to the thermodynamics of hydrogen-atom loss for methyl-substituted heteroaromatic rings.

141

2.5 Conclusions

Using the B3LYP level of theory, the bond dissociation enthalpies for hydrogen

loss and alkyl-group loss were compiled for the alkylated heteroaromatics. Using the

small 6-31G* basis set for geometry optimization, followed by single-point energies

using the larger 6-311+G** basis set, this efficient method was shown to give reliable

results, matching well with experimental values for C-H BDE values when known, and in

agreement with all of the qualitative BDE trends predicted by the CBS-QB3 method.

Overall, loss of hydrogen to form a benzylic-like radical was roughly 10 kcal/mol more favorable than loss of an alkyl group, regardless of ring size or heteroatom. BDE trends varied considerably with respect to ring size and bond type. For the hydrogen-loss reactions, the methyl-substituted azabenzenes saw higher BDEs than their hydrocarbon counterpart, toluene. However, the methyl-substituted five-membered heteroaromatic

rings displayed consistently lower BDEs than the azabenzenes and, moreover, exhibited a

trend where the 2-substituted compounds had lower BDEs than the 3-substituted

compounds, due to increased stabilization of the radical via increased proximity to the

heteroatom. In the hydrogen-atom fragmentation reactions of the ethyl derivatives, the

trends within each class of compounds (azabenzenes vs. heteroaromatic rings) were

duplicated, but an overall stabilization of the radicals of interest caused the BDEs of these

reactions to drop by ~4 kcal/mol. In terms of alkyl-group loss, the ortho-substituted

azabenzenes were more able to stabilize the in-plane sp2 radicals in these reactions, due

to the adjacent nitrogen atom, and had consistently lower BDEs than their counterparts.

The five-membered heteroaromatic rings could not exploit a similar relationship, as their

142

alkyl-loss reactions caused geometric perturbations to the ring, decreasing any aromatic contribution to resonance stability and thus increasing the BDEs for these compounds.

Spin densities were calculated for these heterocyclic radicals and correlated well with the BDE values; this was seen especially in the case of the five-membered heteroaromatics. Temperature effects were explored for the hydrogen-loss reactions and demonstrated that the reactions stayed constant in their endothermicity but became substantially less endoergic over the 298-2000 K range, due to entropic effects on the free energy. The free energy profiles of hydrogen-atom loss compared to alkyl-group loss showed that both pathways experienced similar temperature effects. Finally, the vibrational frequencies attributed to methyl and ethyl rotations were also explored for their effect on the thermodynamic predictions using the hindered rotor approximation, and it was seen that this treatment gave comparable results for the enthalpies and free energies of these reactions.

With respect to predicting the chemistry of the larger heterocyclic systems found in coal, our work suggests that both hydrogen-atom loss and alkyl-group loss reactions will contribute as initiation steps for the high-temperature combustion reactions of these rings. Longer alkyl chains and larger ring sizes will increase reactivity. The initial steps of radical formation are expected to become much more favorable at high temperatures.

This chapter represents the first in a series exploring the combustion pathways of the alkylated heteroaromatic rings. The B3LYP method has shown promise in elucidating these pathways, which have implications for exploring combustion pathways of coal compounds.

143

References for Chapter 2

1 Online. Internet. Available http://www.rst2.edu/ties/acidrain/IEcoal/how.htm

Accessed 31 May 2006.

2 Smith, K. L.; Smoot, L. D.; Fletcher, T. H.; Pugmire, R. J. The Structure and Reaction

Processes of Coal; Plenum Press: New York, 1994.

3 (a) Axworthy, A. Fuel 1987, 57, 29-35. (b) Meyers, R. A. Coal Structure; Academic

Press, New York, 1982.

4 (a) Mackie, J.; Colket, M.; Nelson, P. J. Phys. Chem. 1990, 94, 4099-4106. (b)

Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10339-10348. (c) Kiefer, J.; Zhang, Q.;

Kern, R.; Yao, J.; Jursic, B. J. Phys. Chem. B 1997, 101, 7061-7073.

5 Eisele, F. L. Int. J. Mass Spectrom. Ion Phys. 1983, 54, 119-126.

6 Williams, A.; Pourkashanian, M.; Jones, J.M.; Rowlands, L. J. Inst. Energy, 1997, 70,

102-113.

7 Streit, G. E.; Spall, D. E.; Hall, J. H. Environ. Sci. Technol. 1987, 21, 519-525.

8 (a) Parr, R. G.; Yang, W. Density Functional Theory in Atoms and Molecules; Oxford

University Press: New York, 1989. (b) Labanowski, J. W.; Andzelm, J. Density

Functional Methods in Chemistry; Springer: New York, 1991.

9 Barckholtz, C.; Barckholtz, T. A.; Hadad, C. M. J. Am. Chem. Soc. 1999, 121, 491-500.

10 (a) Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6088-6094; (b) Fadden,

M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6324-6331.

11 Mirvish, S.S. Cancer Lett. 1995, 93, 17-48.

144

12 Terentis, A.; Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10334-10339.

13 Ikeda, E.; Mackie, J. J. Anal. Appl. Pyrolysis 1995, 34, 47-63.

14 Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10339-10348.

15 Frerichs, H.; Schliephake, V.; Tappe, M.; Wagner, H. Z. Phys. Chem. 1990, 166.

16 Yeung, L.; Elrod, M. J. Phys. Chem. A 2003, 107, 4470-4477.

17 Pitz, W. J.; Seiser, R.; Bozzelli, J. W.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin-

Leclerc, F.; Seshadri, K.; Westbrook, C. K. 30th International Symposium on

Combustion; July 2004.

18 (a) Klotz, S. D.; Brezinsky, K.; Glassman, I. Proceedings of the Combustion Institute;

1998, 27. 337-344. (b) Emdee, J. L.; Brezinsky, K.; Glassman, I. J. Phys. Chem. 1992,

96, 2151-2161. (c) Zhong, X.; Bozzelli, J. Int. J. Chem. Kinet. 1997, 29, 893-913. (d)

Zhong, X.; Bozzelli, J. J. Phys. Chem. A 1998, 102, 3537-3555. (e) Linstedt, R. P.

Combust. Sci. Technol., 1996, 120, 119-167. (f) Venkat, C.; Brezinsky, K.; Glassman, I.

Proceedings of the Combustion Institute; 1982, 19, 143-152. (g) Brezinsky, K.; Litzinger,

T.; Glassman, I. Int. J. Chem. Kinet. 1984, 16, 1053-1061. (h) Griffiths, J.; Halford-Maw,

P.; Rose, D. Combust. Flame 1993, 95, 291-306. Roubaud, A.; Minetti, R.; Sochet, L.

Combust. Flame 2000, 121, 535-541.

19 Wen, Z.; Li, Z.; Shang, Z.; Cheng, J.-P. J. Org. Chem. 2001, 66, 1466.

20 Bean, G. P. Tetrahedron 2002, 58, 9941.

21 Nam, P. C.; Nguyen, M. T.; Chandra, A. J. Phys. Chem. A 2005, 109 10342 –10347.

22 Clark, W. D.; Price, S. J. Can. J. Chem. 1970, 48, 1059-1064.

23 Grovenstein, E., Jr.; Mosher, A. J. J. Am. Chem. Soc. 1970, 92, 3810-3812.

145

24 Ebrecht, J.; Hack, W.; Wagner, H. Ber. Bunsenges. Phys. Chem. 1989, 93, 619-623.

25 Ebert, K. H.; Ederer, H. J.; Schmidt, P. S., Baulch, D. L.; Cobos, C. J.; Cox, R. A.;

Esser, C.; Frank, P.; Just, T.; Kerr, J. A.; Pilling, M. J.; Troe, J.; Walker, R. W.; Warnatz,

J. J. Phys. Chem. Ref. Data 1992, 21, 411-429.

26 (a) Baulch, D. L.; Cobos, C. J.; Cox, R. A.; Frank, P.; Hayman, G.; Just, T.; Kerr, J. A.;

Murrells, T.; Pilling, M. J.; Troe, J.; Walker, R. W.; Warnatz, J. J. Phys. Chem. Ref. Data

1994, 23, 847-1033. (b) Scott, M.; Walker, R. W. Combust. Flame 2002, 129, 365-377.

27 Muller-Markgraf, W.; Troe, J. J. Phys. Chem. 1988, 92, 4899 – 4905.

28 (a) Frenklach, M.; Wang, H. Proc. Combust. Inst. 1991, 23, 1559-1556. (b) Wang, H.;

Frenklach, M. J. Phys. Chem. 1994, 98, 11465-11489. (c) Frenklach, M. Phys. Chem.

Chem. Phys. 2002, 4, 2028-2037.

29 (a) Hippler, H.; Troe, J. J. Phys. Chem. 1990, 94, 3803-3806. (b) Walker, J.; Tsang,

W. J. Phys. Chem. 1990, 94, 3324-3327. (b) Ellison, G. B.; Davico, G. E.; Bierbaum, V.

M.; DePuy, C. H. Int. J. Mass Spectrom. Ion Proc. 1996, 156, 109-131; (c) Muralha, V.;

dos Santos, R.; Simoes, J. J. Phys. Chem. A 2004, 108, 936-942.

30 McMillen, D. F.; Golden, D. M. Annu. Rev. Phys. Chem. 1982, 33, 493-532.

31 Song, S.; Golden, D. M.; Hanson, R. K.; Bowman, C. T. J. Phys. Chem. A 2002, 106,

6094.

32 Meot-Ner, M. M.; Lias, S. G. Binding Energies Between Ions and Molecules, and the

Thermochemistry of Cluster Ions. In NIST Chemistry WebBook, NIST Standard Reference

Database Number 69. Linstrom, P. J., Mallard, W. G., Eds. National Institute of

Standards and Technology: Gaithersburg, MD, March 2003.

146

33 Smith, L., Jr.; Evans, S. J. Chem. Soc. Perkin Trans. 2 2000, 7, 1541-1551.

34 Gronert, S. J. Org. Chem. 2006, 71, 1209-1219.

35 Wallington, T. J.; Kaiser, E. W.; Farrell, J. T. Chem. Soc. Rev. 2006, 35, 335-347.

36 Gaussian 98, Revision A.11.03, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria,

G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A., Jr.;

Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K.

N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci,

B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.;

Cui, Q.; Morokuma, K.; Salvador, P.; Dannenberg, J. J.; Malick, D. K.; Rabuck, A. D.;

Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Baboul, A. G.; Stefanov,

B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.;

Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.;

Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Andres, J. L.; Gonzalez, C.; Head-

Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian, Inc., Pittsburgh PA, 2001.

37 Gaussian 03, Revision C.02, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G.

E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.;

Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi,

M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.;

Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.;

Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;

Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;

Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.;

147

Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain,

M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.;

Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.;

Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham,

M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.;

Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A. Gaussian, Inc., Wallingford CT,

2004.

38 (a) Becke, A. D. Phys. Rev. A. 1988, 38, 3098-3100. (b) Lee, C.; Yang., W.; Parr, R.

G. Phys. Rev. B. 1988, 37, 785-789. (c) Becke, A. D. J. Chem. Phys. 1993, 98, 1372-

1377.

39 Bauschlichter, C. W., Jr.; Langhoff, S. R. Mol. Phys. 1999, 96, 471-476.

40 (a) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A., Jr. J. Chem. Phys. 1996,

104, 2598-2619; (b) Petersson, G. A.; Frisch, M. J. J. Phys. Chem. A 2000, 104, 2183-

2190.

41 Scott, A. P.; Radom, L. J. Phys. Chem. 1996, 100, 16502-16513.

42 (a) Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1989, 90, 2154. (b) Gonzalez, C.;

Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523.

43 Reed, A. E.; Weinstock, R. B.; Weinhold, F. J. Chem. Phys. 1985, 83, 735-746.

44 Hehre, W. J.; Radom, L.; Schleyer, P. v. R.; Pople, J. A. Ab Initio Molecular Orbital

Theory; John Wiley and Sons: New York, 1986.

45 Carter, A. Classical and statistical thermodynamics. Upper Saddle River: Prentice

Hall, 2001.

148

46 Using software designed by Dr. Timothy Barckholtz (Exxon-Mobil), obtained via

private communication.

47 Amir,-Ebrahimi, V.; Choplin, A.; Demaison, J.; Roussy, G. J. Mol. Spectros. 1981, 89,

42-52.

48 Bera, P. P.; Schaefer, H. F., III. Proc. Natl. Acad. Sci. USA. 2005, 102, 6698-6703.

149

CHAPTER 3

COMBUSTION PATHWAYS OF THE ALKYLATED HETEROAROMATICS: PEROXY RADICAL PATHWAYS OF THE ALKYLATED AZABENZENES

3.1 Introduction

The structure of coal is notoriously difficult to study with any degree of precision,

given its complex structure and the variation of that structure with geography.1 However,

an understanding of coal’s combustion processes is valuable, given the fact that this fossil

fuel accounts for a great deal of America’s electrical power,2 and has recently shown

promise for use in hydrogen fuel generation.3 A potential approach for examining the

chemistry of coal specimens involves the fact that aromatic hydrocarbons and

heteroaromatic rings are recurring components of these larger structures.4 Since it is

known what types of coal dominate in certain parts of the country, knowing the reactions available to each type of monocyclic unit can ultimately provide clues as to how the overall coal structure in a certain area might react, better predicting the dominant reactions and likely emissions of a given geographical specimen.

Small aromatic heterocyclic units (Figure 3.1) have been used as models for coal

chemistry in several pyrolytic studies; using shock tube studies, Mackie et al. observed

the thermal decompositions of pyridine,5 while Kiefer et al. examined the processes of

150 pyrazine, pyrimidine, and pyridine;6 commonly observed products of pyrolysis involved

ring-scission species such as HCN, HC≡C-C≡N, and acetylene.

N N N N 2 N 2 2

3 3 5 N N 4 4 4 Pyridine Pyridazine Pyrimidine Pyrazine

Figure 3.1. Heteroaromatic rings of interest. These six-membered heteroaromatic rings are commonly classified as the azabenzenes.

Substantially fewer oxidative studies have been completed. Tabares et al. studied the

oxidation of pyridine with O(3P) and noted a decrease in reactivity relative to benzene.7

Alfassi et al. have examined the formation and reactivity of pyridylperoxy radicals in solution.8 Additionally, Eisele’s work on pyridine and picoline ions in the troposphere

suggested that the low-temperature oxidation pathways of these species have implications

for atmospheric processes.9 This supposition was validated by Yeung and Elrod, who

performed chemical ionization mass spectrometric studies of the reactions of hydroxyl

(HO•) radical with pyridine, the picolines, the lutidines, and the ethylpyridines, and

postulated that these pyridinated compounds could indeed have substantial implications

on tropospheric ion content.10 These heteroaromatics have the potential to be oxidized at

151 11 the nitrogen atom and thus form NOx species, which can subsequently react to over-

produce tropospheric ozone and contribute to acid rain.12

Our group has examined several aspects of heteroaromatic combustion.

Barckholtz et al. explored the use of various methods and basis sets in calculating the

C−H bond dissociation enthalpies (BDEs) of a number of unsubstituted

heteroaromatics.13 It was determined that DFT methodology (B3LYP) afforded the most

useful balance of accuracy and computational cost, and this approach has been used in

much of our subsequent work on these compounds. Fadden et al. examined the oxidation

pathways of these species, noting that after addition of molecular oxygen at 298 K, five-

membered heteroaromatic rings were most likely to lose atomic oxygen,14 while the six-

membered heteroaromatics (azabenzenes) favored dissociation back to molecular oxygen

and the relevant aromatic radical.15 Intramolecular pathways competed with these reactions, regardless of ring size, and in some cases had implications for NOx formation.

As temperatures increased, the entropic benefit of dissociative pathways increased. For

the alkyl-substituted heteroaromatic rings, the chemistry of these compounds is likely

complicated via their dual nature: both alkyl and aromatic. Our previous work suggests

that the chemistry of the relevant peroxy radicals affords a logical starting point for

considering these effects.

Studies of the alkylated heteroaromatic rings and their oxidized derivatives are

limited in the literature despite the fact that these species actually provide the most

rational model for coal chemistry. Mackie et al. have postulated several pathways in the

thermolytic decomposition of the picolines,16 while Frerichs et al. examined the reaction

152 of the picolines with atomic oxygen.17 Therefore, it is also instructive to consider the

pathways of the equivalent alkylated aromatic hydrocarbon and oxidized aromatic

hydrocarbon species—in this case, toluene and benzylperoxy radical—as model species.

Several studies have been completed on the oxidation of toluene in general and formation

of benzylperoxy radical in particular. Emdee et al. proposed that toluene’s combustion is

most sensitive to its reaction with oxygen to form benzyl radical and hydroperoxyl

radical. Dagaut et al. proposed that toluene combustion proceeds mainly through

generation of the benzyl radical, which can thermally decompose to acetylene and

cyclopentadienyl or oxidatively decompose to phenyl and formyl radicals (via

benzaldehyde).18 Pitz et al. have generated comprehensive mechanisms for toluene combustion in varying settings, due to its widespread use as a fuel additive.19 Andino et

al. modeled the atmospheric photooxidation of toluene and substituted xylenes, noting

that several cyclized peroxy radical rearrangement products were energetically stable.20

Elmaimouni et al. studied the equilibrium between benzylperoxy radical and its

component parts (benzyl radical and O2), calculating the enthalpy of the peroxy radical

formation pathway to be –20.1 kcal/mol, and extrapolating the free energy to be −11.4

21 22 kcal/mol. Fenter et al. completed a kinetic study which proposed kf of (7.6 ± 2.4) x

-13 3 4 -1 10 exp[(190 ± 160)K/T] cm /molecule-s, Kp of (6.3 ± 0.2) x 10 atm , and an enthalpy

of reaction of –21.8 kcal/mol. Hoyermann and Seeba used a flow reactor coupled to

23 mass spectrometric detection to study the same equilibrium, determining kf = (4.44 ±

11 3 -1 1.3) x 10 cm / molecule-s and Kp = 57,200 bar .

153 Several plausible routes are available to this species, some dissociative and others

intramolecular (Figure 3.2). Benzylperoxy radical can lose atomic oxygen to form benzyloxy radical, dissociate back to molecular oxygen and benzyl radical; dissociate into phenyl radical and the cyclized side chain; cyclize at various positions within the

benzene ring (ipso, ortho, meta, and para); or undergo a hydrogen transfer to yield a

benzylic hydroperoxide radical. Initial calculations on these pathways were used to

generate hypothetical routes for the corresponding azabenzylperoxy species, which were

subjected to a more rigorous study. The experimental data available for benzylperoxy

radical also allowed us to corroborate our computational approach before moving on to

the more pertinent calculations of the azabenzenes.

154 CH2O +O(3P)

O + O A

B O O CH2OO

+O2 C

D H OO

G E F

O O H CHOOH O O

H

Figure 3.2. Likely pathways for the alkylated aromatic peroxy radicals, using benzylperoxy radical as a model compound.

The work outlined in this study modeled the oxidation pathways of the alkylated azabenzenes at 298 K, examining both the kinetic and thermodynamic aspects of these reactions. The picolinylperoxy radicals were studied in greatest detail; geometries and electron distributions were studied to elucidate the changes in reactivity resulting from the nitrogen atom’s presence in the ring. The general trends of temperature effects, increasing alkyl substitution, and varying approximations (harmonic oscillator vs. rotor)

155 were also examined. Additionally, the thermodynamics of the reactions of the

picolinylperoxy radicals were compared to those of the alkylated diazabenzylperoxy

radicals, to explore the effect of multiple nitrogen atoms on these compounds’ reactivity.

A subsequent study will examine similar pathways in several five-membered

heteroaromatic rings (furan, thiophene, pyrrole, and oxazole) that also have relevance to

coal chemistry; these smaller compounds will afford the additional benefit of being able

to observe the effects of varying and multiple heteroatoms, while comparing the results of

the two studies will elucidate any potential effects of ring size.

3.2 Computational Methods

All geometry optimizations, vibrational frequency calculations, and single-point

energy calculations were completed with Gaussian 9824 and 0325 at the Ohio

Supercomputer Center. The B3LYP/6-31G* hybrid density functional theory (DFT)26 level was used for all geometry optimizations and vibrational frequency calculations.

Single-point energies were calculated at the B3LYP/6-311+G** level, using six Cartesian d functions with the scf=tight option; it has been shown that there is a small basis set effect with more polar systems,27 so this larger basis set was selected to compensate for

that effect. Natural Population Analyses (NPA)28 were also obtained at the B3LYP/6-

311+G**//B3LYP/6-31G* level of theory. Additionally, certain geometries and energies

were also obtained using Complete Basis Set (CBS-QB3) methods,29 to validate the DFT

approach for these particular systems.

156 Vibrational frequencies were calculated for each stationary point to characterize

these structures as minima or transition states. The unscaled vibrational frequencies were

used to calculate the thermodynamic corrections to the enthalpy and free energy. Once

obtained, zero-point vibrational energy corrections were scaled by a factor of 0.9806.30

The overall enthalpy at each temperature was determined from the single-point energy, the thermal correction to the enthalpy, and the scaled zero-point energy, while the overall free energy at each temperature also included the entropic correction to the free energy.

Transition states were connected to reactants and products by either using intrinsic reaction coordinate (IRC)31 searches or displacing the relevant geometries by ±

10% along the reaction coordinate characterized by the imaginary vibrational frequency,

then carefully optimizing (opt = calcfc) these resulting geometries to their corresponding

minima.

For the doublet radical species and many transition states, spin contamination

() was negligible (no greater than 0.80). The DFT methods employed in this work have demonstrated an ability to minimize spin contamination for the radicals of interest,

keeping the computed values near the expected value of 0.75. Spin densities were

obtained for the hydrogen-loss radicals using the natural population analysis (NPA)

method.32

Because flame processes are some of the reactions of interest for these

compounds, it is necessary to see how these species react at higher temperatures. The

enthalpies of both the parent compounds and the resultant radicals were calculated from

298 to 2000 K via the temperature-dependent term:33

157 =Δ trans + rot + vib )()()()( + RTTHTHTHTH

3 3 i / kThv −1 )( ++=Δ ∑ i evNhRTRTTH )1( +− RT 2 2 i

The summation takes place over all 3N−6 normal modes. For each compound, the scaled zero-point energy and enthalpy contribution were computed at various temperatures via the given equations and added to the B3LYP total energy, to obtain the bond dissociation enthalpies as a function of temperature. A similar approach was used with calculation of the entropic corrections, to determine the free energies of reaction over the same temperature range.34

To reduce potential errors in the thermodynamic values arising from treating all

vibrations according to the harmonic-oscillator rigid-rotor approximation, low-energy

torsions in the alkyl-substituted species were also treated as hindered rotors,35 and the resulting energetics were compared to those from the original calculations.

3.3 Results

Our first set of calculations examined the kinetics and thermodynamics of the pathways available to benzylperoxy radical, calibrating our computational approach against experimental findings by way of DFT and complete basis set calculations. The most likely pathways were delineated and selected for further exploration with the alkylated heteroaromatic rings.

The oxidation pathways of the azabenzenes— pyridine, pyrimidine, pyridazine, and pyrazine— were then subjected to a thorough analysis. This work began from where our earlier work36 ended: i.e. the most stable radical species were formed from H-atom

158 loss at the benzylic position of the compound. We explored the various conformations of

the peroxy radicals that could result from reaction of these heterocyclic radicals with

molecular oxygen, to determine the minimum-energy conformation, thus allowing

calculation of enthalpies and energies of oxidation. Once these peroxy radical geometries were obtained, the possible pathways available to these species were examined and their enthalpies and energies compiled. High-temperature pathways were also considered.

The nomenclature used throughout the remainder of the discussion will indicate

that all radicals are located on the terminal oxygen of the compound; that is, 2-

picolinylperoxy radical will be used as an abbreviation for 2-picolin-2’-ylperoxy (c-

• C5H4N-CH2OO ) radical; similarly, 2-ethylpyridinylperoxy radical will be used to refer to

• 2-ethylpyridin-2’-ylperoxy radical (c-C4H4N2-C(H)(CH3)OO ). Finally, unless otherwise

specified, the reactions will be discussed in the context of free energies of activation and

free energies of reaction at 298 K.

3.4 Discussion

3.4.1 Toluene as reference compound

As with the case of our previous work on the C−-H and C−alkyl bond dissociation enthalpies of the heteroaromatic rings, more information is available on the hydrocarbon analogue of these species, in this case benzylperoxy radical. These data provide a chance to calibrate our computational method against experimental results, even though directly relevant experimental studies are not yet available. Thus, the

159 oxidation pathways of benzyl radical and subsequent reactions of benzylperoxy radical

were explored via DFT and CBS-QB3 methods. As mentioned in the introduction, the

enthalpy of reaction for the oxidation of benzyl radical has been calculated to be –20.1

and –21.8 kcal/mol in two different studies;21,22 the free energy of reaction has also been

extrapolated to be –11.4 kcal/mol. The CBS calculations yielded results within 1

kcal/mol of experiment, predicting an enthalpy of oxidation of –22.7 kcal/mol, with a free

energy of oxidation of –12.2 kcal/mol. Moreover, both DFT and CBS methods predicted

comparable results for the possible benzylperoxy radical pathways (Table 3.1):

cyclization to form a five-membered ring was most favorable, followed by cyclization to form a four-membered ring, followed by an internal 1,3-hydrogen-atom transfer that ultimately yields benzaldehyde and hydroxyl radical.

160

B3LYP CBS-QB3 Experiment ΔΗ ο ΔG ο ΔΗ ο ΔG ο ΔΗ ο ΔG ο Dissociation to -16.0 -5.5 -22.7 -12.2 -20.1a -12.2 benzyl radical -21.8b + O2

B3LYP CBS-QB3 ΔΗ ο ΔG ο ΔΗ ο ΔG ο Path A 56.5 47.3 61.9 52.5 Path B 53.7 42.9 55.8 45.3 Path C 33.2 34.2 25.8 26.9 Path D 35.6 23.7 14.6 15.1 Path E NO CYCLIZATION Path F NO CYCLIZATION Path G -33.9 -35.4 -37.7

B3LYP CBS-QB3 ΔΗ ‡ ΔG‡ ΔΗ ‡ ΔG‡ Path C 35.7 37.2 30.1 31.7 Path D 33.5 35.6 29.7 31.8 Path G 37.9 38.9 38.4 39.2

Table 3.1. Comparison of DFT (B3LYP/6-311+G**//B3LYP/6-31G*) and CBS-QB3 results for key processes in the reactions possible for benzylperoxy radical. (See Figure 3.2 for the different pathways.)

DFT replicated the qualitative trends obtained by the CBS-QB3 calculations, which correspondingly replicated the quantitative values obtained by experiment. The DFT energies therefore provided an economic route to valid energetic conclusions and will be referenced in the course of this discussion.

At 298 K, the likeliest pathways included cyclization at the ortho position to form a five-membered ring, with a barrier of 35.6 kcal/mol (Pathway D); cyclization at the ipso

161 formation to form a four-membered ring, with a barrier of 37.2 kcal/mol (Pathway C);

and internal H-transfer, with a barrier of 38.9 kcal/mol (Pathway G). This last step was

particularly interesting. The internal hydrogen transfer facilitated by the sp3 C−H bonds of the benzylic position was expected to yield a benzylic hydroperoxide radical; however, after displacement of the relevant transition state and minimization to products, it was seen that this transition state connected the benzylperoxy radical directly to benzaldehyde and hydroxyl radical. The dissociative nature of the reaction and stability of the products formed led to a highly favorable free energy of reaction, which would be expected to become moreso with increasing temperature. An analogous reaction would not be possible for phenylperoxy radical and the related azabenzylperoxy radicals; it is the additional methylene unit of benzylperoxy radical that facilitates this unique path, generating a stable carbonyl species as well as the reactive hydroxyl radical (thereby contributing to the radical pool). Geometries for cyclization at the meta and para positions failed to optimize; this was unsurprising, with the amount of strain these species would incur given their starting aromatic geometries. Dissociative processes were endoergic at low temperatures, as expected.

At 298 K, the rearrangements available to the benzylperoxy radical dictate the

low-temperature combustion pathways, and the kinetic trends largely mimic the

thermodynamic trends. It was assumed, however, that the dissociative pathways would

become more likely at temperatures increased into the high-temperature combustion

range. Six pathways (dissociation to reactants, ipso cyclization, ortho cyclization, H-

162 transfer, oxygen loss, and dioxirane loss) were thus explored for each of the

picolinylperoxy radicals (Figure 3.3).

O N H OO O

3.3

3.4a

OO O

N N CH 2OO N N CH 2O +O2

3.4b 3.2 3.1

N CH OO H N CHO N HO O + + O 3.5

N CH 2O +O(3P) 3.7

3.6

Figure 3.3. Numbering scheme for the azabenzylperoxy radicals.

3.4.2 Methyl-substituted pyridines (Picolines)

As with benzylperoxy radical, the reactions of 2-picolinylperoxy radical were first studied via DFT methods, calibrating the most likely pathways against CBS-QB3 energies (Table 3.2). Overall, the B3LYP method predicts reaction energies that are actually 6-7 kcal/mol lower than those predicted by CBS-QB3. Most importantly, it was

163 again seen that DFT and CBS matched qualitative trends; thus, the introduction of a

heteroatom did not affect our computational approach.

B3LYP CBS-QB3 ΔΗ ΔG ΔΗ ΔG 3.2 → 3.1 16.3 6.5 22.7 13.2 TS(3.2-3.3) 37.1 39.4 31.0 33.4 3.2 → 3.3 34.2 35.7 26.1 27.9 TS(3.2-3.4a) 30.4 33.2 25.6 28.7 3.2 → 3.4a 20.6 23.2 13.2 16.0 TS(3.2-3.4b) 47.1 49.8 3.2 → 3.4b -5.8 -4.1 TS(3.2-3.5) 38.0 39.4 37.6 39.2 3.2 → 3.5 -24.7 -33.6 -23.7 -32.4 3.2 → 3.6 61.6 53.0 3.2 → 3.7 49.2 38.7

Table 3.2. Comparison of reaction pathways available to 2-picolinylperoxy radical at 298 K, via B3LYP/6-311+G**//B3LYP/6-31G*. All enthalpies and energies are relative to the peroxy radical (2); when preceded by TS, the relative data are the enthalpy and energies of activation. Where applicable, 4a refers to cyclization at carbon and 4b refers to ortho cyclization at nitrogen. For the most likely pathways, B3LYP results were calibrated against CBS-QB3 calculations.

The numbering scheme used in Figure 3.3 will be adopted over the course of the paper, as

each alkylated heterocyclic peroxy radical (3.2) is systematically examined: the routes of interest include dissociation back to reactants (3.1); cyclization at the ipso position (3.3);

cyclization at the ortho position, which could be either carbon (3.4a) or nitrogen (3.4b)37;

internal H-transfer and subsequent scission (3.5); loss of atomic oxygen (3.6); and loss of

dioxirane (3.7). Where relevant, transition states will be referred to as “TS” between the

164 two numbered species; i.e., the transition state for cyclization at the ipso position would

be TS(3.2-3.3).

Comparing the 2-picolinylperoxy radical pathways to those of benzylperoxy radical illuminated interesting trends. Overall, the same pathways were preferred for both 2-picolinylperoxy radical and benzylperoxy radical. Cyclization at the ortho position (3.2 → 3.4a) yielded a stable five-membered ring, while cyclization at the ipso position (3.2 → 3.3) yielded a four-membered ring; the strain present in the four-

membered ring disfavored ipso cyclization relative to ortho cyclization. Ortho

cyclization at nitrogen was also a plausible route to yield a stable nitrosyl radical, but this

occurred with a substantially higher barrier. The internal H-transfer yielded hydroxyl

radical and 2-picolinal, presumably via a benzylic-like radical; the barrier for this

pathway was roughly equivalent to that of ipso cyclization. Thus, ipso cyclization, ortho

cyclization (at carbon), and H-transfer followed by scission are the processes likely for 2-

picolylperoxy radical in low-temperature combustion (atmospheric) reactions.

In terms of comparative quantitative energies of the reactions, loss of dioxirane to

form the 2-pyridyl radical (3.2 → 3.7) was less endoergic than in the hydrocarbon

analogue; this makes sense in the context of our previous work in this area, which has

suggested that the nitrogen atom in 2-pyridinyl radical increases its resonance stability,

rendering it more stable than phenyl radical. Aside from this one exception, the

picolinylperoxy radical pathways were more endoergic than the corresponding benzylperoxy radical pathways; several aspects of the relevant structures were examined

to discern the cause of this effect.

165 As can be seen in Figure 3.4, the introduction of the nitrogen atom introduces some slight changes into the aromatic ring.

Cmeth O2

Cmeth C6 O1 N1 C6 O2 C1 C5 C2 O1

C2 C5 C C3 4 C4 C3

Benzylperoxy radical 2-Picolinylperoxy radical Change in bond length C6-C1 1.402 N1-C2 1.341 -0.061 C1-C2 1.400 C2-C3 1.399 -0.001 C2-C3 1.395 C3-C4 1.393 -0.002 C3-C4 1.396 C4-C5 1.395 -0.001 C4-C5 1.397 C5-C6 1.395 -0.002 C5-C6 1.394 C6-N1 1.338 -0.052 C1-Cmeth 1.502 C2-Cmeth 1.508 0.006 Cmeth-O1 1.476 Cmeth-O1 1.462 -0.014 O1-O2 1.320 O1-O2 1.322 0.002

Figure 3.4. Change in bond lengths with addition of nitrogen atom to the aromatic system. All bond lengths in angstroms and obtained from the B3LYP/6-31G* optimized structures.

Most notably, both carbon-nitrogen bonds are shortened by more than 0.5 angstroms, due to the increased electronegativity difference. Overall, however, the ring structures are highly similar. The difference in preferred dihedral orientation is apparent between the hydrocarbon and heteroaromatic ring; it seems possible that however,

166 rotation of the relevant dihedral angles occurs with small barrier heights (< 1 kcal/mol), such that this could not account for any substantive differences. In terms of the aromaticity lost in the course of the cyclization reactions, benzene has a resonance energy of 36 kcal/mol,38 while pyridine’s resonance energy is 34 kcal/mol.39 It seems that the cyclization steps, each of which involves loss of aromaticity, would thus be less favorable for benzylperoxy than for picolinylperoxy, contrary to the calculated energetics.

One possible answer to these energetic differences might be that the allylic system remaining in the cyclized hydrocarbon peroxy radicals more readily stabilizes the resultant radical than does that of the heterocyclic derivatives. The excess spin densities of a representative cyclization were examined for both benzylperoxy and 2- picolinylperoxy radicals (Figure 3.5). The resulting allylic system for both ortho- cyclized species involves atoms 2, 1, 6, 5, and 4, as numbered. It can be seen that the two cyclized derivatives experience virtually identical spin densities, with notable increases in spin density on the terminal and central atoms of the allylic system (2, 6, 4) and corresponding decreases on the other two atoms (1, 5). The only discernable difference between the two systems arises on atom 1, which is carbon in the benzylperoxy derivative and nitrogen in the picolinylperoxy derivative; the nitrogen experiences substantially less of a change (roughly half compared to carbon) in its excess spin density. It can be hypothesized that the presence of the nitrogen might slightly disrupt the allylic delocalization in the heterocyclic derivatives and thus account for the higher energies of reaction seen for the alkylated azabenzylperoxy radicals.

167

1 1 2 7 N 2 7 6 6 O 8 O 8 3 3 5 O 5 O 9 9 4 H 4 H

Position Benzylperoxy radical 2-Picolinylperoxy radical 1 -0.11 -0.06 2 0.23 0.22 3 0.09 0.10 4 0.21 0.20 5 -0.09 -0.09 6 0.26 0.25 7 -0.01 -0.01 8 -0.22 -0.21 9 -0.40 -0.40

Figure 3.5. Variation in excess spin density via cyclization at the ortho position, for benzylperoxy radical and 2-picolinylperoxy radical. Excess spin density (α−β) calculated via comparison between cyclized product and relevant peroxy radical, using NPA analyses (B3LYP/6-311+G**//B3LYP/6-31G*). Numbering scheme used for convenience of reference.

Another possible factor involves the negative inductive effect of the nitrogen’s presence in the aromatic ring. Taberes et al. noted a smaller rate constant for the reaction of pyridine with O(3P) as compared to benzene and O(3P), attributed to this phenomenon.

Although the referenced work discussed the aromatic rings as pure , it is plausible that a related trend could be repeated in the radical pathways of these related compounds, hence the relative kinetic barriers of benzylperoxy radical and its nitrogen analogue.

168 The reaction enthalpies and energies are summarized for 3-picolinylperoxy radical and 4-picolinylperoxy radical (Table 3.3). The barrier to ortho cyclization was consistently lowest, roughly 2-4 kcal/mol less endoergic than the barrier to ipso cyclization, and 5−6 kcal/mol lower than the barrier to hydrogen transfer. 3-

Picolinylperoxy radical presented an interesting case in that both of its ortho cyclization pathways (3.2 → 3.4a and 3.2 → 3.4b) generated a new carbon-carbon bond; neither ortho cyclization occurred at nitrogen. Between these two options, cyclization adjacent to nitrogen was slightly more favorable than at the position para to nitrogen. 3-

Picolinylperoxy and 4-picolinylperoxy radicals did allow for a more careful delineation of the hydrogen-atom transfer pathway. The values reported in Table 3.3 reflect complete dissociation to products (for the benefit of comparison to previous data), but actual displacement of the transition state in these two cases resulted in a stable complex and a benzylic-like radical, respectively.

169

3-picolinylperoxy radical ΔΗ ΔG 3.2 → 3.1 15.6 5.6 TS(3.2-3.3) 36.2 38.2 3.2 → 3.3 33.5 34.8 TS(3.2-3.4a) 32.8 35.4 3.2 → 3.4a 21.3 23.8 TS(3.2-3.4b) 34.0 36.6 3.2 → 3.4b 24.4 26.8 TS(3.2-3.5) 38.6 39.8 3.2 → 3.5 -29.3 -38.3 3.2 → 3.6 56.3 49.3 3.2 → 3.7 53.8 43.2

4-picolinylperoxy radical ΔΗ ΔG 3.2 → 3.1 15.2 5.6 TS(3.2-3.3) 38.3 40.6 3.2 → 3.3 35.7 37.5 TS(3.2-3.4) 31.8 34.7 3.2 → 3.4 21.9 24.5 TS(3.2-3.5) 38.7 40.1 3.2 → 3.5 -28.7 -37.4 3.2 → 3.6 62.7 53.8 3.2 → 3.7 52.0 41.6

Table 3.3. Compiled enthalpies and energies of reaction (kcal/mol) for 3- and 4- picolinylperoxy radical (B3LYP/6-311+G**//B3LYP/6-31G*). Where applicable, 4a refers to cyclization at carbon and 4b refers to ortho cyclization at nitrogen. All energies and enthalpies relative to those of the relevant peroxy radical (2).

170 C O N O C C + C C

C

0.0 39.4 -33.6

N

C C

C C

C C

O

O

0.0 39.8 -34.3

N

C C

C C

C

C

O

O

0.0 40.1 0.1

Figure 3.6. H-transfer pathways delineated for each of the three picolinylperoxy radicals. Free energies of reaction at 298 K calculated at the B3LYP/6- 311+G**//B3LYP/6-31G* level of theory.

Simultaneously looking at all three picolinylperoxy radicals (Figure 3.6) allows a better understanding of the overall reaction. As seen for 4-picolinylperoxy radical, the hydrogen-atom transfer yields a hydroperoxide-functionalized, benzylic-like radical, with a free energy of reaction of only 0.3 kcal/mol. This had been the expected product in all

171 three cases; however, the 4-substituted derivative was the only one to yield the actual

radical. For 3-picolinylperoxy radical, displacement of the transition state and minimization to products yielded a complex of 3-picolinal and hydroxyl radical, presumably stabilized by an apparent hydrogen bond between the carbonyl oxygen and the hydroxyl hydrogen; the free energy of reaction for this step was –34.3 kcal/mol, while the infinitely separated 3-picolinal and hydroxyl radical were at –38.3 kcal/mol relative to the initial peroxy radical. The 2-picolinylperoxy radical’s hydrogen-atom transfer was included as a frame of reference; displacement from to the transition state yielded separated products, with a free energy of –33.6 kcal/mol. This unusual reactivity is facilitated by the dual nature of the alkylated heterocyclic peroxy radical. The 1,3-H transfer is of interest given its ability to ultimately form a stable carbonyl species and reactive hydroxyl radical, and will be studied further.

Overall, the picolinylperoxy radicals all demonstrate the kinetic preferences of

ortho cyclization, ipso cyclization, and internal H-transfer over all potential pathways at

298 K. While NOx formation did not seem probable at 298 K due to the high barrier of

the cyclization at nitrogen, the relevant N−O radicals were thermodynamically stable and merited further attention at higher combustion temperatures.

3.4.3 Temperature effects for the picolinylperoxy radicals

As the temperature of combustion increases, dissociative pathways become vastly

more favorable due to entropic effects. These effects can be monitored by calculating the

thermal corrections to enthalpy and free energy at a variety of temperatures. Graphs are

172 provided for the case of 2-picolinylperoxy radical (Figures 3.7 and 3.8). In terms of

thermodynamics, raising the combustion temperature generally increased the enthalpy of

reaction slightly for all pathways considered. As expected, the free energies of reaction

changed dramatically, with the dissociative pathways becoming vastly more exoergic.

This was most evident in the case of atomic oxygen loss (3.2 → 3.6) and dioxirane loss

(3.2 → 3.7), as both these pathways were substantially endoergic at 298 K. The peroxy radical cyclizations generally saw a smaller decrease in free energy of reaction; ortho cyclization at carbon was still preferred over ipso cyclization at carbon at 2000 K.

Notably, both the processes of ortho cyclization at the nitrogen atom (3.2 → 3.4b) and of internal H-transfer followed by scission to hydroxyl and 2-picolinal (3.2 → 3.5) were exoergic at 298 K, and became substantially moreso with increasing temperature.

173

Figure 3.7. Variation in ΔHrxn with temperature for the pathways available to 2- picolinylperoxy radical. 3.2 → 3.3 denoted by open diamond; 3.2 → 3.4a denoted by solid square; 3.2 → 3.4b denoted by open triangle; 3.2 → 3.5 denoted by x; 3.2 → 3.6 denoted by dash; 3.2 → 3.7 denoted by solid diamond. All enthalpies calculated at the B3LYP/6-311+G**//B3LYP/6-31G* level of theory.

174

Figure 3.8. Variation in ΔGrxn with temperature for the pathways available to 2- picolinylperoxy radical. 3.2 → 3.3 denoted by open diamond; 3.2 → 3.4a denoted by solid square; 3.2 → 3.4b denoted by open triangle; 3.2 → 3.5 denoted by x; 3.2 → 3.6 denoted by dash; 3.2 → 3.7 denoted by solid diamond. All energies calculated at the B3LYP/6-311+G**//B3LYP/6-31G* level of theory.

The effects of temperature on barrier heights were also explored for 2-

picolinylperoxy radical (Figures 3.9 and 3.10). The lowest barriers at 298 K involve the peroxy radical processes: cyclizations (TS(3.2-3.3), TS(3.2-3.4a), and TS(3.2-3.4b)) and internal H-transfer (TS(3.2-3.5)). The activation barriers for these four routes were explored at temperatures up to 2000 K. While the activation enthalpies rose slightly over the given temperature range for all four routes, maintaining the relative preferences, the activation energies differed. While activation energies steadily decreased in all four

175 cases, the effect was substantially more pronounced for TS(3.2-3.5). This pathway

competed with TS(3.2-3.3) at 298 K for the second most favorable barrier (TS(3.2-3.4a)

being dominant); near 1250 K, TS(3.2-3.5) overtook all of the other pathways, due to its

greater entropic effects, and became the major high-temperature decomposition pathway.

The H-transfer pathway is unique in that it proceeds through a peroxy radical

rearrangement to yield two products; it exhibits characteristics of both low-temperature

and high-temperature oxidation reactions. Thus, it is not surprising that this reaction

competes across temperature ranges. Notably, TS(3.2-3.4b) continuously demonstrated a

relatively high activation barrier in both low and high temperatures. This is the pathway

that demonstrates the greatest implication for NOx formation; however, it seems as though this hazard would be circumvented at all but the highest combustion temperatures.

176

Figure 3.9. Change in ΔHactivation with increasing temperature, for 2-picolinylperoxy radical. Activation enthalpy for 3.2 → 3.3 denoted by open diamond; activation enthalpy for 3.2 → 3.4a denoted by solid square; activation enthalpy for 3.2 → 3.4b denoted by open triangle; activation enthalpy for 3.2 → 3.5 denoted by x. All enthalpies calculated at the B3LYP/6-311+G**//B3LYP/6-31G* level of theory.

177

Figure 3.10. Change in ΔGactivation with increasing temperature, for 2-picolinylperoxy radical. Activation energy for 3.2 → 3.3 denoted by open diamond; activation enthalpy for 3.2 → 3.4a denoted by solid square; activation enthalpy for 3.2 → 3.4b denoted by open triangle; activation enthalpy for 3.2 → 3.5 denoted by x. All energies calculated at the B3LYP/6-311+G**//B3LYP/6-31G* level of theory.

The enthalpies and energies for the pathways of the 3- and 4-substituted

picolinylperoxy radicals were compiled over varying temperatures (Tables 3.4 and 3.5).

Similar trends were seen for these derivatives; TS(3.2-3.4a) consistently demonstrated the lowest reaction barrier at 298 K, while the dissociative nature of TS(3.2-3.5) made its reaction barrier lowest at 2000 K. Each of these pathways is clearly only the first step in the combustion of these species; subsequent possible steps will be discussed.

178

3-picolinylperoxy radical 3.2 → 3.3 3.2 → 3.4a 3.2 → 3.4b 3.2 → 3.5 3.2 → 3.6 3.2 → 3.7 298 22.5 34.7 25.6 -27.9 57.9 52.4 500 23.7 36.1 26.8 -26.4 59.5 52.3 750 25.4 38.0 28.6 -25.1 61.2 51.5 1000 27.1 40.0 30.3 -24.0 62.9 48.7 1250 28.8 41.5 31.9 -23.0 64.4 49.8 1500 30.4 43.1 33.6 -22.1 65.8 48.9 1750 32.0 44.7 35.1 -21.3 67.2 47.9 2000 33.6 46.3 36.7 -20.5 68.6 47.0

4-picolinylperoxy radical 3.2 → 3.3 3.2 → 3.4 3.2 → 3.5 3.2 → 3.6 3.2 → 3.7 298 35.7 22.0 -28.5 62.6 52.1 500 35.9 22.0 -28.3 62.9 51.6 750 36.1 22.1 -28.5 63.1 50.9 1000 36.2 22.2 -29.0 63.1 50.1 1250 36.4 22.4 -29.6 63.1 49.2 1500 36.5 22.5 -30.3 62.9 48.3 1750 36.6 22.5 -30.9 62.9 47.3 2000 36.7 22.6 -31.7 62.7 46.4

Table 3.4. Variation in enthalpies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*) for 3- and 4-substituted picolinylperoxy radicals, with increasing temperature.

179 3-picolinylperoxy radical 3.2 → 3.3 3.2 → 3.4a 3.2 → 3.4b 3.2 → 3.5 3.2 → 3.6 3.2 → 3.7 298 32.6 21.5 24.5 -40.4 45.7 41.8 500 30.7 20.5 23.5 -49.3 37.1 35.0 750 27.6 18.5 21.1 -61.0 25.4 26.9 1000 23.9 16.0 18.8 -73.1 13.2 19 1250 19.8 13.0 15.7 -85.5 0.7 11.4 1500 15.3 9.7 12.3 -98.1 -12.2 3.9 1750 10.5 6.2 8.6 -110.8 -25.3 -3.8 2000 5.5 2.4 4.7 -123.7 -38.7 -10.6

4-picolinylperoxy radical 3.2 → 3.3 3.2 → 3.4 3.2 → 3.5 3.2 → 3.6 3.2 → 3.7 298 37.6 24.6 -37.2 53.9 41.8 500 38.8 26.3 -43.2 48.0 35.0 750 40.2 28.4 -50.6 40.4 26.9 1000 45.2 30.6 -57.9 32.9 19.1 1250 42.8 32.7 -65.0 25.4 11.4 1500 44.1 32.7 -72.0 17.8 3.9 1750 45.4 36.7 -78.9 10.3 -3.4 2000 46.6 38.7 -85.8 2.8 -10.6

Table 3.5. Variation in free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*) for 3- and 4-substituted picolinylperoxy radicals, with increasing temperature.

3.4.4 Ethyl-substituted pyridines (thermodynamics):

The ethyl-substituted pyridinylperoxy radicals were also studied, to gauge the

effect of increasing alkyl substitution on these species’ reactivity (Table 3.6). Overall,

the free energies of both activation and reaction were comparable to those of the methyl- substituted picolines. The dissociative pathway 3.2 → 3.1 saw a larger ΔGo for the

methyl-substituted pyridinylperoxy radicals. This makes sense given the greater stability

that accompanies increasing substitution on an alkyl radical center; the ethyl radicals

180 would be more stable than their methyl counterparts, so dissociation to these more stable

species would be more favorable. The ipso cyclization pathway 3.2 → 3.3 demonstrated

a lower barrier height and a lower energy of reaction for the ethyl derivatives, as did the

H-atom transfer/scission pathway 3.2 → 3.5.

2-Substituted 3-Substituted 4-Substituted Methyl Ethyl Methyl Ethyl Methyl Ethyl 3.1 6.5 5.1 5.6 4.9 5.6 4.2 TS(3.2-3.3) 39.4 37.7 38.2 36.1 40.6 38.4 3.3 35.7 34.1 34.8 32.6 37.5 35.3 TS(3.2-3.4a) 33.2 34.7 35.4 34.2 34.7 33.0 3.4a 23.2 23.8 23.8 22.3 24.5 23.1 TS(3.2-3.4b) 49.8 48.9 36.6 37.2 34.7 33.0 3.4b -4.1 1.8 26.8 27.3 24.5 23.1 TS(3.2-3.5) 39.4 37.8 39.8 38.4 40.1 38.8 3.5 -33.6 -35.7 -38.3 -41.0 -37.4 -40.9 3.6 53.0 48.3 49.3 49.7 53.8 48.7 3.7 38.7 32.4 43.2 36.1 41.6 34.8

Table 3.6. Comparison of free energy profiles (kcal/mol) at 298 K for the picolinylperoxy (methylpyridinylperoxy) and ethylpyridinylperoxy radicals. Where applicable, 4a refers to cyclization at carbon and 4b refers to ortho cyclization at nitrogen. All energies relative to those of the relevant peroxy radical (2) and calculated at the B3LYP/6-311+G**//B3LYP/6-31G* level of theory. For each pair of structures, the more favorable energetic barrier or energy of reaction is indicated by bold italics, to best illustrate the differences in reactivity caused by increasing the length of the alkyl chain.

These results were counterintuitive, since it would seem that the ethyl-substituted

compounds would experience more steric strain in the cyclization steps. The additional

radical stabilization afforded to the ethyl derivatives outweighs these steric effects;

moreover, given the planar nature of both the initial aromatic ring and the newly-formed

181 ring, the additional carbon of the methyl group does not come near the new bond, so the

actual steric strain sustained by the reaction is less than would be imagined.

Ortho cyclization afforded the most varied results. For the 2-substituted

derivatives, the ethyl-substituted pyridinylperoxy radical saw a higher barrier to

cyclization and higher energy of reaction, at carbon (TS(3.2-3.4a)) this might be attributed in part to the greater van der Waal’s forces present in the ethyl derivative, which could have destabilizing interactions with the nitrogen. Interestingly, 2- ethylpyridinylperoxy radical showed a slightly lower barrier to cyclization at nitrogen

(TS(3.2-3.4b)), even while 2-methylpyridinylperoxy radical had a substantially lower energy of reaction for this process. The 3-substituted derivatives likewise exhibited contradictory results. For cyclization at C-2, the ethyl derivative reacted more readily; for cyclization at C-4, the methyl derivative did. This would seem to support our earlier hypothesis that cyclization at C-4 incurs a degree of geometric strain for the 3-substituted derivatives that is not seen in the other ortho cyclization pathway. While in the 3- methylpyridinylperoxy radical, the difference in energy between the two cyclizations was roughly 1.2 kcal/mol, in the 3-ethylpyridinylperoxy radical, this difference was more than

twice as high (3.0 kcal/mol). The additional alkyl group present in the ethyl derivative

increases the strain in the less favorable TS(3.2-3.4b) transition state. Finally, for the 4-

substituted derivatives, 4-ethylpyridinylperoxy radical reacted more readily and formed a

more stable ortho cyclization derivative (the 3.4a and 3.4b pathways were equivalent in

each case due to symmetry) than did 4-methylpyridinylperoxy radical; both the distance

182 from nitrogen and the relative planarity of the newly-formed ring likely contributed to

these effects.

Loss of atomic oxygen (3.2 → 3.6) was approximately 5 kcal/mol more favorable for the 2- and 4-ethyl-substituted derivatives than their methyl counterparts, due to the increased alkyl stabilization of the ethyl oxyradical over the methyl oxyradical.

In the case of substitution at the 3-position, the methyl and ethyl derivatives saw nearly identical free energies of reaction. Our group has previously noted that the unusual reactivity of 3-substituted pyridinyl derivatives, 3-pyridinyloxy radical in particular, and it is plausible that the stability afforded to 3-methylpyridinyloxy radical stems from similar factors. Finally, loss of the side chain (3.2 → 3.7) to form the aromatic radical and cyclized dioxiranyl species was roughly 6 kcal/mol more favorable for the ethyl derivatives regardless of substitution, which must be attributed to the increased stability of methyl dioxirane over dioxirane, since the resulting aromatic radicals are identical.

It is worth reiterating that, in the most likely pathways at 298 K, the differences in energetics were generally less than 2 kcal/mol between the methyl and ethyl derivatives.

Moreover, the trends of the ethyl derivatives echo those of the methyl derivatives: ortho cyclization was the most favorable pathway, followed by ipso cyclization and hydrogen- atom transfer. Thus, the methyl derivatives provide a good qualitative guide for the overall reactions of alkylated heteroaromatic rings, even while the quantitative energies may vary slightly.

183 3.4.5 Diazabenzene trends

From the exhaustive study of the picolinylperoxy radicals, it is evident that of the

cyclization pathways available, two main cyclizations, at the positions ipso and ortho to

the alkyl substitution, will dominate the chemistry at 298 K. Moreover, an internal H-

transfer will compete with both of these cyclization possibilities at 298 K. In certain

cases, a less-kinetically-favorable ortho cyclization pathway yields a stable N−O radical that increases the thermodynamic benefits of this reaction. At higher temperatures, the

dissociative processes of oxyradical formation, aromatic radical formation will see an

entropic benefit that will increase their energetic favorability; the aforementioned H-

transfer ultimately leads to a carbonyl-functionalized aromatic radical and hydroxyl

radical, so is a factor in both low and high-temperature combustion.

The same six pathways (dissociation, ipso cyclization, ortho cyclization, atomic

oxygen loss, dioxirane loss, and internal H-transfer) were considered as the major

combustion pathways for the azabenzylperoxy radicals. The free energies at 298 K for

these pathways were compiled (Table 3.7). Comparing the thermodynamics of these

reactions to those of the picolinylperoxy radicals, it is evident that the second nitrogen

atom does not alter the qualitative trends of reaction; ortho cyclization at carbon still

demonstrates the lowest kinetic barrier, followed by ipso cyclization and hydrogen-atom

transfer. While these two-nitrogen species would seem to double the potential for NOx formation seen in the picolinylperoxy radicals, the barrier of such a process is still prohibitive.

184

Parent Diazabenzene Substitution 3.1 3.3 3.4a 3.4b 3.5 3.6 3.7 Pyridazine 3 6.0 36.5 22.2 -9.1 -32.4 50.5 39.5 4 4.1 35.1 24.7 24.0 -36.8 50.5 39.8 Pyrimidine 2 6.3 37.5 0.2a 0.2b -33.6 52.5 40.4 4 5.8 39.8 25.3 1.1 -31.5 50.9 38.4 5 8.2 33.9 26.2 26.2b -35.2 51.6 45.6 Pyrazine 2 4.4 33.3 21.3 -6.1 -34.1 50.4 37.5 a For 2-methylpyrimidinylperoxy radical, both 4a and 4b refer to ortho cyclization at nitrogen. bOrtho pathways equivalent via symmetry.

Table 3.7. Free energies of reaction at 298 K (kcal/mol) for reaction pathways available to the peroxy radicals of the diazabenzenes. Where applicable, 3.4a refers to cyclization at carbon and 3.4b refers to ortho cyclization at nitrogen. All reaction pathways relative to the peroxy radical (3.2) of the relevant parent compound.

Overall, the picolinylperoxy radicals provide excellent predictive models for the

diazabenzenes; moreover, these peroxy radicals also demonstrate fairly consistent

quantitative energetics. Kinetic barriers and temperature profiles for some reactions of

interest of the diazabenzenes are included in Supporting Information and compare well to the overall energetics of the picolinylperoxy radicals.

3.4.6 Approximations

Calculations were run to determine the validity of the harmonic oscillator

approximation in obtaining the thermal corrections to the enthalpy and entropy (Table

3.8). The lowest-energy vibrations (i.e., the motions of the side chain) were identified

and treated as rotations, using a hindered rotor approximation rather than the harmonic

oscillator. The free energies of reaction reflected small changes in most cases, but these

185 were generally less than 2 kcal/mol. Considering that this variation in approximation

would have the most impact on the entropy of these species, it seems logical that the free energies would see more of an effect than the enthalpies (Supporting Information).

However, these slight quantitative differences did not yield any difference in the relative favorability of the pathways available to the peroxy radicals.

Parent Azabenzene Substitution 3.1 3.3 3.4a 3.4b 3.5 3.6 3.7 Pyridine 2 7.6 36.8 24.3 -3.0 -32.5 53.5 39.9 3 7.1 36.3 25.3 28.3 -36.8 48.4 44.6 4 7.3 37.5 26.0 26.0a -36.0 55.2 43.1 Pyridazine 3 6.7 37.3 22.9 -8.4 -31.7 46.6 40.2 4 6.6 36.7 26.2 25.5 -35.2 48.0 41.3 Pyrimidine 2 7.5 38.7 1.4 1.4a -32.4 48.8 41.5 4 6.5 39.3 24.8 0.6 -32.0 52.6 38.0 5 5.6 35.3 27.6 27.6a -36.7 48.2 44.1 Pyrazine 2 5.2 34.1 22.2 -5.3 -33.2 52.1 38.3 aOrtho cyclization identical via symmetry.

Table 3.8. Free energies of reaction at 298 K (kcal/mol), using rotor approximation in place of harmonic oscillator approximation in calculating the thermal correction to the enthalpy and entropy. Energies calculated at the B3LYP/6-311+G**//B3LYP/6-31G* level of theory.

3.4.7 Subsequent steps

Clearly, the peroxy radical pathways delineated in this article are only the first of

many steps involved in the combustion of these complicated compounds. Overall, logical subsequent pathways can be delineated for the products of several of the peroxy radical pathways discussed, many of which lead into already-established combustion routes.

186 As shown (Figure 3.11), aromatic radicals can be generated either directly or gradually via several of the intermediates discussed above. Following hydrogen-atom transfer and scission (3.5), the resulting 2-picolinal can dissociate into 2-pyridinyl radical and formyl radical. The oxyradical (3.6) can lose its side chain to form 2-pyridinyl radical and formaldehyde; likewise, loss of dioxirane (3.7) directly generates 2-pyridinyl radical.

These aromatic radicals can either cleave the aromatic ring (at high temperatures) to form multiple smaller products or be oxidized at the incipient radical center; these pyrolytic processes are fairly prevalent in the literature, and the oxidative decomposition processes of the non-alkylated aromatic radicals have been thoroughly studied by our group.

Correspondingly, ipso cyclization (3.3) could be followed by re-opening of the four- membered ring to replace the original Cring−Cside_chain bond with a Cring−O bond, then losing formaldehyde to generate the non-alkylated oxyradical. This latter species has again been modeled via previous work in our group.14,15 While the overall decompositions of these species are quite complex, the reactions delineated herein can be viewed as a subset of the chemistry available to these compounds.

187 O N N N O OOCH2 O 3.3 +CH2O

N CHOOH N CHO N O + 3.5 H N CH2OO

+ HO

3.2

N CH2O N

+CH2O 3.6

+O(3P)

N O + O

3.7

Figure 3.11. Subsequent steps available to certain peroxy radical rearrangement products, which could build on extant combustion pathways known for the non-alkylated azabenzyl radicals and oxy radicals.

3.5 Conclusions

Coal combustion has enormous energetic implications, and yet this process is chemically difficult to understand, given the complexity and variation of coal structure.

188 One approach to simplifying this problem is to model the oxidation processes of smaller

aromatic hydrocarbons and heteroaromatic rings, then extrapolate from these smaller

compounds to better understand coal chemistry. In this work, we have delineated key

steps available to the oxidized derivatives of alkylated azabenzenes, mainly using free

energies (298 K) of activation and reaction obtained at the B3LYP/6-

311+G**//B3LYP/6-31G* level of theory.

Benzylperoxy radical was first explored to obtain a qualitative picture of the likely peroxy radical pathways for a species containing both alkyl and aromatic components. From these initial calculations, it was determined that cyclizations at the ipso and ortho positions, relative to side chain substitution, were most likely at 298 K.

Also competitive was an internal 1,3-H transfer that could occur via the side chain of the species, ultimately yielding benzaldehyde and hydroxyl radical. These three pathways were modeled, along with dissociation back to the benzylic-like radical and molecular oxygen, loss of atomic oxygen, and loss of dioxirane; these latter three processes were expected to contribute heavily at high temperatures.

The picolinylperoxy radicals (methylpyridinylperoxy radicals) were explored at

298 K. It was seen that their chemistry mimicked that of benzylperoxy radical: ortho

cyclization at carbon occurred with a barrier of around 30 kcal/mol; ipso cyclization

occurred with a barrier of around 35-40 kcal/mol; internal H-atom transfer occurred with

a barrier of roughly 40 kcal/mol. The additional functionality afforded by the nitrogen

atom allowed a cyclization resulting in a stable N−O radical, which has the potential for

NOx chemistry; however, this process had a substantially high barrier (~50 kcal/mol) and

189 was unlikely to be a factor at atmospheric temperatures. The energetics of the 2-

picolinylperoxy radical were closely compared to those of benzylperoxy radical, since the only difference between them was a nitrogen atom. It was seen that while the qualitative trends were replicated between benzylperoxy radical and its nitrogen analogue, 2- picolinylperoxy radical had slightly higher activation barriers and energies of reaction in nearly every case, which was attributed to the relative stabilization of the allylic systems in the cyclized derivatives of the two species. Additionally, CBS-QB3 calculations were performed on both benzylperoxy and 2-picolinylperoxy radical and demonstrated that the

DFT approach performed well in replicating the trends seen via the CBS method.

Temperature effects were explored; over the range 298 to 2000 K, the inherently

dissociative processes (atomic oxygen loss, molecular oxygen loss, dioxirane loss) saw a

large entropic benefit and became substantially exoergic. The rearrangement pathways

decreased somewhat in energy, as did the kinetic barriers to these processes. Notably, the

hydrogen-atom transfer pathway saw the most precipitous drop in its barrier and became

the dominant processes around 1250 K. This reaction was of considerable interest given

its relevance to both low-temperature and high-temperature combustion, as a

rearrangement pathway that culminates in dissociation to a stable aldehyde-

functionalized heteroaromatic compound and hydroxyl radical. Even at higher

temperatures, it was seen that the reaction resulting in N−O radical formation occurred

with a barrier substantially higher than that of other rearrangements; this process likely

only contributes minimally and at extremely high flame temperatures to NOx formation.

190 The effects of increasing alkyl substitution were explored. It was seen that the same rearrangement pathways were favored for the ethylpyridinylperoxy radicals as for their methyl counterparts. Generally, the energies of activation and reaction for the most likely pathways were comparable, within 2 kcal/mol of one another, which suggests that the methyl compounds can provide good models for the alkylated heteroaromatic rings in general. Small differences in reaction energetics were attributed to inductive effects of the nitrogen atom and to differing amounts of geometric strain introduced via cyclization, where applicable.

After the picolinylperoxy radicals had been thoroughly modeled, these results were compared to the energies of reaction for the relevant processes of the alkylated diazabenzenes (pyridazine, pyrimidine, and pyrazine). It was seen that the presence of a second nitrogen atom had little effect on either the identity or the energetics of the preferred pathways.

Finally, due to the presence of low-energy torsions in the alkylated peroxy side chain, the effect of changing from a harmonic oscillator approximation to a rotor approximation in the case of these lowest vibrations was explored for all the methylazabenzylperoxy radicals. It was seen that this change did slightly affect the thermal corrections to the enthalpy and entropy and was most evident in the free energies of reaction. However, these changes were small (< 2 kcal/mol) and did not affect the relative preferences of the pathways.

In summary, these findings suggest that the overall chemistry of the alkylated heterocyclic peroxy radicals is consistent regardless of alkyl substitution or number of

191 nitrogen atoms. The picolinylperoxy radicals provide excellent models for the chemistry exhibited by this larger class of species. Moreover, aromatic hydrocarbons can themselves predict several aspects of this chemistry, the exception being those processes involving oxidation or rearrangement at the nitrogen atom itself. Subsequent work in this area will explore comparable pathways for the five-membered heteroaromatic rings, in an attempt to discern the effects of both ring size and of different heteroatoms on the favorability of these reactions. It is hoped that these studies of models for coal chemistry will elucidate key combustion processes of this important energy source.

192 References for Chapter 3

1 Smith, K. L.; Smoot, L. D.; Fletcher, T. H.; Pugmire, R. J. The Structure and Reaction

Processes of Coal; Plenum Press: New York, 1994.

2 Online. Internet. Available http://www.rst2.edu/ties/acidrain/IEcoal/how.htm

Accessed 31 May 2006.

3 “University Bolsters Major Endeavor to Solve Global Energy Problems,” Ohio State

University News in Engineering. Online. Internet. Available

http://engineering.osu.edu/nie/article.php?e=783&s=3&a=2. Accessed 15 February

2007.

4 (a) Axworthy, A. Fuel 1987, 57, 29-35. (b) Meyers, R. A. Coal Structure; Academic

Press, New York, 1982.

5 Mackie, J.; Colket, M.; Nelson, P. J. Phys. Chem. 1990, 94, 4099-4106.

6 Kiefer, J.; Zhang, Q.; Kern, R.; Yao, J.; Jursic, B. J. Phys. Chem. B 1997, 101, 7061-

7073.

7 Taberes, F.L.; Gonzalez-Urena, A. J. Phys. Chem. 1983, 87, 4933-4936.

8 Alfassi, Z. B.; Khaikin, G. I.; Neta, P. J. Phys. Chem. 1995, 99, 4544-4548.

9 Eisele, F. L. Int. J. Mass Spectrom. Ion Phys. 1983, 54, 119-126.

10 Yeung, L.Y.; Elrod, M. J. J. Phys. Chem. A 2003, 107, 4470-4477.

11 Williams, A.; Pourkashanian, M.; Jones, J.M.; Rowlands, L. J. Inst. Energy, 1997, 70,

102-113.

12 Streit, G. E.; Spall, D. E.; Hall, J. H. Environ. Sci. Technol. 1987, 21, 519-525.

193

13 Barckholtz, C.; Barckholtz, T. A.; Hadad, C. M. J. Am. Chem. Soc. 1999, 121, 491-

500.

14 Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6324-6331.

15 Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6088-6094.

16 Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10339-10348.

17 Frerichs, H.; Schliephake, V.; Tappe, M.; Wagner, H. Z. Phys. Chem. 1990, 166.

18 Dagaut, P.; Pengloan, G.; Ristori, A. Phys. Chem. Chem. Phys. 2002, 4, 1846-1854.

19 (a) Pitz, W. J.; Seiser, R.; Bozzelli, J.W.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin-

Leclerc, F.; Seshadri, K.; Westbrook, C.K. 2nd Joint Meeting of the U.S. Section of the

Combustion Institute: Oakland, CA. March 25-28, 2001. (b) Pitz, W.J.; Seiser, R.;

Bozzelli, J.W.; Seshadri, K.; Chen, C. J.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin-

Leclerc, F.; Westbrook, C. K. Thirtieth Symposium (International) on Combustion. The

Combustion Institute; Chicago, IL: July 25-30, 2004.

20 Andino, J. M.; Smith, J. N.; Flagan, R. C.; Goddard, III, W. A.; Seinfield, J. H. J. Phys.

Chem.1996, 100, 10967-10980.

21 Elmaimouni, L.; Minetti, R.; Sawerysyn, J. P.; Devolder, P. Int. J. Chem. Kinet. 1993,

25, 399-413.

22 Fenter, F.F.; Noziere, B.; Caralp, F.; Lesclaux, R. Int. J . Chem. Kinet. 1994, 26, 171-

189.

23 Hoyermann, K.; Seeba, J. 25th Symp. (Int’l) on Combustion; The Combustion Institute,

1994; 851-858.

194

24 Gaussian 98, Revision A.11.03, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria,

G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A., Jr.;

Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K.

N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci,

B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.;

Cui, Q.; Morokuma, K.; Salvador, P.; Dannenberg, J. J.; Malick, D. K.; Rabuck, A. D.;

Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Baboul, A. G.; Stefanov,

B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.;

Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.;

Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Andres, J. L.; Gonzalez, C.; Head-

Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian, Inc., Pittsburgh PA, 2001.

25 Gaussian 03, Revision C.02, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G.

E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.;

Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi,

M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.;

Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.;

Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;

Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;

Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.;

Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain,

M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.;

Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.;

195

Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham,

M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.;

Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A. Gaussian, Inc., Wallingford CT,

2004.

26 (a) Becke, A. D. Phys. Rev. A. 1988, 38, 3098-3100. (b) Lee, C.; Yang., W.; Parr, R.

G. Phys. Rev. B. 1988, 37, 785-789. (c) Becke, A. D. J. Chem. Phys. 1993, 98, 1372-

1377.

27 Bauschlichter, C. W., Jr.; Langhoff, S. R. Mol. Phys. 1999, 96, 471-476.

28 (a) Reed, A. E.; Weinstock, R. B.; Weinhold, F. J. Chem. Phys. 1985, 83, 735-746. (b)

Reed, A. E; Curtiss, L. A.; Weinhold, F. Chem. Rev. 1988, 88, 899-926.

29 (a) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A., Jr. J. Chem. Phys. 1996,

104, 2598-2619; (b) Petersson, G. A.; Frisch, M. J. J. Phys. Chem. A 2000, 104, 2183-

2190.

30 Scott, A. P.; Radom, L. J. Phys. Chem. 1996, 100, 16502-16513.

31 (a) Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1989, 90, 2154. (b) Gonzalez, C.;

Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523.

32 Reed, A. E.; Weinstock, R. B.; Weinhold, F. J. Chem. Phys. 1985, 83, 735-746.

33 Hehre, W. J.; Radom, L.; Schleyer, P. v. R.; Pople, J. A. Ab Initio Molecular Orbital

Theory; John Wiley and Sons: New York, 1986.

34 Carter, A. Classical and statistical thermodynamics. Upper Saddle River: Prentice

Hall, 2001.

196

35 Using software designed by Dr. Timothy Barckholtz (Exxon-Mobil), obtained via private communication.

36 Chapter 2 of this work.

37 In the cases where the side chain is not adjacent to a nitrogen atom (i.e. 3- picolinylperoxy radical), pathway 4b also reflects cyclization at carbon.

38 Carey, F. A. Organic Chemistry, 5th Edition. New York: McGraw-Hill, 2003.

39 Wiberg, K. B.; Nakaji, D.; Breneman, C. M. J. Am. Chem. Soc. 1989, 111, 4178-4190.

197

CHAPTER 4

COMBUSTION PATHWAYS OF THE ALKYLATED HETEROAROMATICS: PEROXY RADICAL PATHWAYS OF THE FIVE-MEMBERED HETEROAROMATICS

4.1. Introduction

Coal is a ubiquitous energy source, accounting for 50% of the United States’

electricity, and this prevalence is likely to increase in the future, given coal’s

inexpensiveness and abundance. Despite its relevance as an energy source, the

combustion chemistry of coal has not been as fully explored as that of other fuels, due to

its complex and varying structure.1 One potential approach to remedying this deficit

would be to obtain a better understanding of the chemistry of the repeating aromatic (both

hydrocarbon and heterocyclic) units that are present in coal; as studies have shown how

coal’s structure varies with geography, understanding how the more prevalent units of the

structure would react could facilitate predictions of which reactions would be more likely

for a given coal sample.

Small heteroaromatic units have been used as models for coal chemistry in several

pyrolytic studies. The azabenzenes have historically been used most often: using shock

tube studies, Mackie et al. observed the thermal decompositions of pyridine,2 while

Kiefer et al. examined the processes of pyrazine, pyrimidine, and pyridine;3 commonly

198

observed products of pyrolysis involved ring-scission species such as HCN, N≡C−C≡CH,

and acetylene. Also of interest are the five-membered heteroaromatics: furan, oxazole,

pyrrole, and thiophene (Figure 4.1).

H N O S O 5 2 2 2 2 4 3 3 3 N Pyrrole Furan Thiophene Oxazole

Figure 4.1. Common five-membered heteroaromatic rings and their numbering schemes.

These species have been previously studied in various capacities (not solely as models for coal chemistry). Besides their role in coal combustion, these have been implicated as products of biomass burning,4 residential fires,5 waste tire burning,6 cigarettes,7 and motor vehicle emissions.8 Bruinsma et al. examined the pyrolysis of the

heteroaromatics most common in coal volatiles,9 determining a rank of relative stability:

furan < cyclopentadiene < pyrrole < pyridine < benzene < thiophene; moreover, they

noted that an additional aromatic ring stabilized most derivatives. Cullis and Norris

similarly studied the pyrolysis of some heteroaromatics, finding that methane and

benzene were the major products, and that the heteroatoms were generally not present in

the major combustion products, being lost as water, , or hydrogen

cyanide.10 Braslavsky and Heicklen compiled a review of the thermal and photochemical

decomposition of some heteroaromatic rings containing nitrogen, oxygen, and sulfur.11

199

Klein et al. studied the variation in heterocyclic bond dissociation enthalpies in

substituted aromatic compounds.12

In particular, furan and pyrrole, as well as their methylated derivatives, have been

common targets of pyrolytic studies. Pyrolysis of furan and its methylated derivatives

has been shown to lead to several products, including carbon monoxide, acetylene,

acetaldehyde, propyne, and allene.13 Pyrrole likewise leads to a variety of products,

including hydrogen cyanide, propyne, allene, and acetylene, cis-crotonitrile, and allyl

cyanide; it has been hypothesized to be pyrolyzed primarily through formation of a

biradical that can break down or recyclize.14

Despite these compounds’ relevance to combustion chemistry, however, their

oxidative decompositions have not been comprehensively explored. Our group has sought to delineate these important pathways. Barckholtz et al. completed an exhaustive survey to select an appropriate computational method for study of the heteroaromatic rings; density functional theory (DFT)15 was shown to balance accurate results with computational economy to the greatest extent.16 Via comparisons of bond dissociation

enthalpies in monocyclic and polycyclic heteroaromatic rings, this study also allowed the

determination of a crucial conclusion: that the monocyclic heteroaromatics do provide

constructive models for the larger systems present in coal. Our group then turned to exploring the oxidations of these model compounds. Fadden et al. saw that there were considerable differences between the pathways of the five-membered heteroaromatics’ peroxy radicals and the six-membered azabenzenes’ peroxy radicals;17 the former could

lose atomic oxygen at a relatively low energetic cost, while for the six-membered rings,

200

losing atomic oxygen was substantially more unfavorable, to the extent that dissociation

back to the aromatic radical and molecular oxygen was preferable. For both sets of

compounds, intramolecular cyclizations competed with oxygen loss, and some of these

cyclizations did lead to nitroso compounds with NOx implications.

Besides strictly aryl units, coal structures also include alkyl units and alkyl-

substituted aromatic rings. Thus, we have recently studied the initial bond dissociation enthalpies and energies of a variety of alkylated heteroaromatic rings.18 With respect to

predicting the chemistry of the larger heterocyclic systems found in coal, these

calculations suggested that both the hydrogen-atom and alkyl-group loss reactions will

contribute as initiation steps for the high-temperature combustion reactions of these rings.

Moreover, longer alkyl chains will increase reactivity, and the azabenzene units are more

likely to react than the five-membered heterocyclic rings. The initial steps of radical

formation are expected to become much more energetically favorable at high

temperatures. Additionally, in a related study, we have explored the oxidative pathways

of the alkylated azabenzenes;19 our work suggested that the overall chemistry of the

alkylated azabenzyl peroxy radicals is consistent regardless of alkyl substitution or

number of nitrogen atoms. Moreover, the picolinylperoxy radicals provided excellent

models for the chemistry exhibited by this larger class of species, and aromatic

hydrocarbons can themselves predict several aspects of this chemistry, the exception

being those processes involving oxidation or rearrangement at the nitrogen atom itself.

This chapter is intended as to supplement these earlier works, examining the

oxidative pathways of the radicals generated from the alkylated five-membered

201

heteroaromatic rings. These findings will allow conclusions on how reactivity varies

with a given heteroatom, as well as with ring size; additionally, the structure of oxazole

structure is such that the effect of multiple heteroatoms can also be observed. It will be

instructive to observe how the additional methylene unit present in these species affects

the overall oxidative pathways, to gain an understanding of whether the alkyl or aromatic

nature of these species dictates the chemistry.

It is well known that preferred oxidation pathways differ between low and high

temperatures. For peroxy radicals, rearrangements and hydrogen-atom transfers are

common steps at atmospheric temperatures. Given the stronger C−H bonds present in aromatic species, these H-atom transfers become less prevalent. Conversely, rearrangements facilitate cyclizations in these species, generating bicyclic derivatives that can further decompose to products of environmental interest. An additional aspect of heterocyclic chemistry involves the potential for NOx and SOx formation, via oxidation of

the heteroatoms present in the molecules; NOx and SOx are notorious pollutants

responsible for such processes as ozone over-production and acid rain formation in the

troposphere.20 As the peroxy radicals arising from pyrrole and thiophene will have both a relevant heteroatom and an oxidation source in close proximity, some of their reactions

could have implications for these emissions, and will be examined as such.

The rearrangements, hydrogen-atom transfers, and other pathways of interest to

low-temperature combustion of the alkylated heteroaromatic peroxy radicals have been

explored. This study focused on the pathways available to the five-membered

heteroaromatic rings. Herein, we report findings on the energies of activation and

202

reaction for these species, noting implications to atmospheric chemistry and high-

temperature combustion processes.

4.2. Computational Methods

All geometry optimizations, vibrational frequency calculations, and single-point

energy calculations were completed with Gaussian 9821 and 0322 at the Ohio

Supercomputer Center. The B3LYP/6-31G* hybrid density functional theory (DFT)23 level was used for all geometry optimizations and vibrational frequency calculations.

Single-point energies were calculated at the B3LYP/6-311+G** level, using six Cartesian d functions with the scf=tight option; it has been shown that there is a small basis set effect with more polar systems,24 so this larger basis set was selected to compensate for

that effect. Additionally, certain geometries and energies were also obtained using

Complete Basis Set (CBS-QB3) methods,25 to validate the DFT approach for these

particular systems.

Vibrational frequencies were calculated for each stationary point to characterize

these structures as minima or transition states. The unscaled vibrational frequencies were

used to calculate the thermodynamic corrections to the enthalpy and free energy. Once

obtained, zero-point vibrational energy corrections were scaled by a factor of 0.9806.26

The overall enthalpy at each temperature was determined from the single-point energy, the thermal correction to the enthalpy, and the scaled zero-point energy, while the overall free energy at each temperature also included the entropic correction to the free energy.

203

Transition states were connected to reactants and products by either using

intrinsic reaction coordinate (IRC)27 searches or displacing the relevant geometries by

±10% along the reaction coordinate characterized by the imaginary vibrational frequency, then carefully optimizing (opt = calcfc) those resulting geometries to their corresponding minima.

For the doublet radical species and many transition states, spin contamination

() was negligible (no greater than 0.80). The DFT methods employed in this work have demonstrated an ability to minimize spin contamination for the radicals of interest,

keeping the computed values near the expected value of 0.75. Spin densities were

obtained using the natural population analysis (NPA) method.28

Because flame processes are some of the reactions of interest for these

compounds, it is necessary to see how these species react at higher temperatures. The

enthalpies of both the parent compounds and the resultant radicals were calculated from

298 to 2000 K via the temperature-dependent term:29

=Δ trans + rot + vib )()()()( + RTTHTHTHTH

3 3 i / kThv −1 )( ++=Δ ∑ i evNhRTRTTH )1( +− RT 2 2 i

The summation takes place over all 3N-6 normal modes. For each compound, the scaled

zero-point energy and enthalpy contribution were computed at various temperatures via the given equations and added to the B3LYP total energy, to obtain the bond dissociation enthalpies as a function of temperature. A similar approach was used with calculation of the entropic corrections, to determine the free energies of reaction over the same temperature range.30

204

To reduce potential errors in the thermodynamic values arising from treating all

vibrations according to the harmonic-oscillator rigid-rotor approximation, low-energy

torsions in the alkyl-substituted species were also treated as hindered rotors,31 and the resulting energetics were compared to those from the original calculations.

4.3. Results and Discussion

We have previously shown the six pathways that are likeliest for these alkylated heteroaromatic peroxy radicals (Figure 4.2). A given peroxy radical may revert back to its benzylic-like parent radical and molecular oxygen (4.2 → 4.1), cyclize at the ipso position (4.2 → 4.3), cyclize at either ortho position (4.2 → 4.4), undergo a 1,3-H transfer followed by scission to an aromatic aldehyde derivative and hydroxyl radical (4.2

→ 4.5), lose atomic oxygen (4.2 → 4.6) or lose dioxirane (4.2 → 4.7). Our related chapter on this subject details the selection of these rearrangements, and also validates our use of the DFT approach at the B3LYP/6-311+G**//B3LYP/6-31G* level to examine the energetics of the pathways.19 Due to the additional heteroatoms present in these compounds, additional pathways will be considered on a case-by-case basis.

205

X O

O

4.3 X

4.4

O O H

X X X CH2 CH2OO CHO +OH +O2 4.1 4.2 4.5

X

3 CH2O +O(P)

4.6

X

+ O O

4.7

Figure 4.2. Rearrangement pathways available to a typical peroxy radical derivative (4.2) of an alkylated five-membered heteroaromatic ring.

The nomenclature used throughout the remainder of the discussion will indicate that all radicals are located on the terminal oxygen of the compound; that is, 2- methylpyrrolylperoxy radical will be used as an abbreviation for 2-methylpyrrol-2’-

• ylperoxy (c-C4H3N-CH2OO ) radical. Ethyl peroxy derivatives are reported as well. In

206

these compounds, the oxygen unit is attached to the benzylic-like carbon, and the radical center again on the terminal oxygen. Thus, 2-ethylpyrrolylperoxy radical will be used as an abbreviation for 2-ethylpyrrol-1’-ylperoxy radical. Furthermore, unless otherwise specified, the peroxy radical pathways of these species will be discussed in terms of their free energies of activation and reaction at 298 K.

Reaction pathways at 298 K were compiled for the methyl-substituted

heteroaromatic peroxy radicals, including activation barriers for the most likely

pathways. Temperature effects were calculated. Since the peroxy radicals and

oxyradicals of interest in this study can adopt multiple conformations, these geometries

were subjected to both intuitive searches and full dihedral scans to discern the true

minima; the peroxy radical and oxyradical geometries reported in this chapter correspond

to the global minima.

Our previous work on the peroxy radicals of the non-alkylated five-membered

heteroaromatic rings demonstrated that oxygen-atom loss was the most accessible and

energetically favored decomposition step at all temperatures, due to the aromatic nature

of the resultant oxyradical. Although the length of the peroxy chain made

rearrangements of these radicals feasible, at both the ipso and ortho positions relative to

peroxy substitution, these rearrangements were generally disfavored relative to oxygen

loss.32 With this current work, it was hypothesized that the results would diverge given

the additional methylene unit between the heteroaromatic unit and the oxygen atoms,

which would prevent complete aromatic delocalization in any of the species of interest

(either the peroxy radicals or their decomposition products).

207

We employed a systematic approach in this related study: first, comprehensively

studying the pathways of the furan derivatives; then, examining the effect of a second

heteroatom on the overall pathways via a study of the oxazole derivatives. Both the

kinetic and thermodynamic behaviors of these species were modeled at temperatures

ranging from 298 to 2000 K. Additionally, the alkylated peroxy radical derivatives of

pyrrole and thiophene were examined, via their kinetic and thermodynamic behavior at

298 K, as well as the temperature profiles of their thermodynamic behavior over a range of combustion temperatures. Overall trends were also explored: the effect of increasing alkyl substitution was examined via the thermodynamics of the related ethyl derivatives’ reactions, and the validity of the harmonic oscillator approximation was discerned via a rigid rotor comparison.

4.3.1. Furan derivatives. The energetics of the 2- and 3-methylfuranylperoxy radicals are

shown (Figures 4.3 and 4.4, Tables 4.1 and 4.2). Reversion to reactants demonstrated a

low energy of reaction but will not be discussed as fully, since we are primarily interested

in those pathways that lead to oxidative decomposition.

208

4.3

O O 4.4 O O

20.1 O H O 27.1 30.2 35.2

4.1 4.2 4.5 O O O 38.8 CH CHO +OH 2 +O2 CH2OO 1.3 0.0 -38.5

4.6 O +O(3P) CH2O 4.7 O 57.6 + O O 52.7

Figure 4.3. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 2-methylfuranylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

209

O O H 4.3 4.4a O O O 33.6 13.8 O O 4.4b 28.9 H 36.2 O 36.3 31.6 4.1 4.2 O O O 39.2 4.5 +O2 O +OH 0.0 3.0 CH2OO CH2 -36.2 CHO

O 4.6 +O(3P) 4.7 O O + O CH2O 48.3 52.4

Figure 4.4. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 3-methylfuranylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

210

298 K 4.1 4.3 4.4a 4.4b 4.5 4.6 4.7 benzylperoxy 5.5 26.9 15.1 15.1 -37.7 52.5 45.3 2-methylfuranylperoxy 1.3 23.9 30.8 a -38.5 48.7 54.3 3-methylfuranylperoxy 3.0 33.6 11.6 31.6 -36.2 48.3 52.4 2-methyloxazolylperoxy 1.8 21.2 7.7 a -34.2 50.9 54.3 4-methyloxazolylperoxy 1.1 32.1 17.5 7.9 -35.2 50.7 47.6 5-methyloxazolylperoxy 5.6 21.1 24.7 a -37.7 53.0 48.3 2-methylpyrrolylperoxy 3.5 25.1 29.1 10.7 -42.7 47.1 52.2 3-methylpyrrolylperoxy 8.0 35.4 21.1 33.0 -39.3 49.3 51.5 2-methylthiophenylperoxy 4.2 22.9 24.2 14.1 -29.3 47.6 48.0 3-methylthiophenylperoxy 5.6 32.8 15.9 27.4 -39.1 47.8 45.3

500 K 2-methylfuranylperoxy -3.1 22.8 30.4 a -43.0 47.8 47.7 3-methylfuranylperoxy -0.1 34.1 12.1 32.7 -44.8 46.0 43.4 2-methyloxazolylperoxy -9.0 20.3 3.7 a -44.8 44.0 40 4-methyloxazolylperoxy -8.9 29.2 15.5 4.5 -45.0 41.5 40.4 5-methyloxazolylperoxy -5.1 15.0 19.3 a -48.0 41.6 42.6 2-methylpyrrolylperoxy 1.9 25.5 29.9 10 -49.2 40.8 44.2 3-methylpyrrolylperoxy -4.8 30.0 13.1 32.7 -51.3 41.1 39.8 2-methylthiophenylperoxy -5.4 23.7 25.6 13.9 -46.1 44.6 41.5 3-methylthiophenylperoxy 0.7 33.4 17.1 28.5 -45.3 43.7 38.9

750 K 2-methylfuranylperoxy -8.7 20.9 29.5 a -54.0 39.6 35.9 3-methylfuranylperoxy -2.9 34.5 12.6 34.0 -53.1 41.6 34.6 2-methyloxazolylperoxy -14.9 21.0 1.3 a -55.3 33.7 28.6 4-methyloxazolylperoxy -14.9 27.2 14.6 1.8 -55.9 34.3 29.0 5-methyloxazolylperoxy -11.1 12.6 17.9 a -58.9 33.7 31.1 2-methylpyrrolylperoxy -1.2 26.0 31.0 8.8 -57.5 32.7 34.2 3-methylpyrrolylperoxy -12.9 25.4 14.2 32.2 -64.0 31.9 26.3 2-methylthiophenylperoxy -8.5 24.5 27.4 13.4 -54.2 39.2 32.5 3-methylthiophenylperoxy -1.4 34.1 18.7 29.9 -53.5 37.8 29.9 aCyclization at this position did not result in a reasonable geometry.

Table 4.1. Free energies of reaction (kcal/mol) with increasing temperature, as calculated at the B3LYP/6-311+G**//B3LYP/6-31G* level of theory. Table 4.1 continued on page 207.

211

Table 4.1, continued.

1000 K 4.1 4.3 4.4a 4.4b 4.5 4.6 4.7 2-methylfuranylperoxy -23.8 18.4 28.0 a -65.3 31.3 23.8 3-methylfuranylperoxy -14.6 34.9 13.0 35.3 -61.1 37.7 25.9 2-methyloxazolylperoxy -30.2 21.6 -1.3 a -66.0 23.2 17.4 4-methyloxazolylperoxy -30.5 24.7 13.4 -1.3 -67.0 27.3 17.4 5-methyloxazolylperoxy -27.0 9.5 15.8 a -70.2 25.5 19.1 2-methylpyrrolylperoxy -13.7 26.3 32.1 7.6 -65.7 24.5 24.5 3-methylpyrrolylperoxy -31.8 20.1 10.6 32.2 -77.3 22.4 12.2 2-methylthiophenylperoxy -21.7 25.3 29.2 12.9 -62.2 34.1 23.8 3-methylthiophenylperoxy -12.8 34.7 20.2 31.2 -61.6 32.2 21.3

1250 K 2-methylfuranylperoxy -40.1 15.5 26.1 a -77 22.9 11.6 3-methylfuranylperoxy -27.2 35.3 13.3 36.6 -69.1 34.3 17.6 2-methyloxazolylperoxy -46.5 22.2 -4.0 a -76.7 12.5 6.2 4-methyloxazolylperoxy -47.3 21.9 11.9 -4.8 -78.2 20.5 5.8 5-methyloxazolylperoxy -41.2 8.9 16.3 a -78.8 20.3 10.1 2-methylpyrrolylperoxy -27.2 26.7 33.1 6.3 -73.7 25.7 5.7 3-methylpyrrolylperoxy -52.9 14.2 6.6 32.2 -91.0 12.8 -2.2 2-methylthiophenylperoxy -35.8 26.1 31.0 12.3 -70.1 29.3 15.4 3-methylthiophenylperoxy -24.9 35.2 21.8 32.6 -69.5 26.9 12.8

1500 K 2-methylfuranylperoxy -44.6 12.2 23.9 a -88.9 14.3 -0.7 3-methylfuranylperoxy -27.5 35.6 13.7 37.9 -77.0 31.1 9.4 2-methyloxazolylperoxy -50.8 22.8 -6.9 a -87.4 1.7 -4.8 4-methyloxazolylperoxy -52.1 19.0 10.2 -8.4 -89.4 13.7 -5.8 5-methyloxazolylperoxy -49.1 2.3 10.8 a -93.2 9.3 -4.7 2-methylpyrrolylperoxy -27.8 27.0 34.2 5.0 -81.6 8.2 1.2 3-methylpyrrolylperoxy -59.9 7.8 2.1 32.2 -105.0 3.0 -16.8 2-methylthiophenylperoxy -36.9 26.9 32.8 11.6 -77.8 24.7 7.1 3-methylthiophenylperoxy -24.5 35.8 23.3 33.9 -77.3 21.8 4.5 aCyclization at this position did not result in a reasonable geometry.

Table 4.1. Continued on page 208.

212

Table 4.1, continued.

1750 K 4.1 4.3 4.4a 4.4b 4.5 4.6 4.7 2-methylfuranylperoxy -61.7 8.6 21.4 a -100.8 5.7 -13.3 3-methylfuranylperoxy -40.3 36 14 39.1 -84.8 28.2 1.3 2-methyloxazolylperoxy -67.6 23.4 -9.9 a -98 -9.1 -15.8 4-methyloxazolylperoxy -69.6 15.8 8.3 -12.3 -100.9 7.1 -17.4 5-methyloxazolylperoxy -66.8 -1.6 8 a -104.8 1.2 -16.6 2-methylpyrrolylperoxy -41.5 27.3 35.2 3.6 -89.4 0.1 -3.4 3-methylpyrrolylperoxy -81.7 1.1 -2.8 32.2 -119.3 -6.8 -31.5 2-methylthiophenylperoxy -51.4 27.6 34.5 11 -84.5 20.2 -1.1 3-methylthiophenylperoxy -36.9 36.3 24.9 29.6 -85 16.9 -3.6

2000 K 2-methylfuranylperoxy -65.3 4.8 18.6 a -113 -3.2 -26 3-methylfuranylperoxy -39.2 36.3 14.3 40.3 -92.4 25.5 -6.6 2-methyloxazolylperoxy -70.9 23.9 -12.9 a -108.6 -20 -26.8 4-methyloxazolylperoxy -73.7 12.5 6.3 -16.2 -112.3 0.6 -29 5-methyloxazolylperoxy -71.1 -5.6 5 a -116.4 -6.8 -28.6 2-methylpyrrolylperoxy -40.7 27.6 36.2 2.3 -97 -8 -12.4 3-methylpyrrolylperoxy -88.6 -5.9 -8 32.2 -133.7 -16.7 -46.5 2-methylthiophenylperoxy -51.3 28.4 36.3 10.4 -93 16 -9.1 3-methylthiophenylperoxy -35.2 36.8 26.4 36.5 -92.6 12.2 -11.6 aCyclization at this position did not result in a reasonable geometry.

213

TS(4.2-4.3) TS(4.2-4.4a) TS(4.2-4.4b) TS(4.2-4.5) 298.15 K 2-methylfuranylperoxy 33.9 39 A 42.3 3-methylfuranylperoxy 36.2 29 36.3 39.4 2-methyloxazolylperoxy 31.2 70.3 A 40.4 4-methyloxazolylperoxy 35 28.1 60 38.3 5-methyloxazolylperoxy 29.5 32.9 A 38.1 500 K 2-methylfuranylperoxy 33.1 38.7 A 40.7 3-methylfuranylperoxy 37.1 30.5 37.8 39.8 2-methyloxazolylperoxy 30.5 70 A 39 4-methyloxazolylperoxy 34.2 27.9 59.2 36.9 5-methyloxazolylperoxy 28.3 32.3 A 36.3 750 K 2-methylfuranylperoxy 31.6 37.9 A 38.3 3-methylfuranylperoxy 38.4 32.6 39.9 40.3 2-methyloxazolylperoxy 29.6 69.5 A 37.1 4-methyloxazolylperoxy 32.9 27.3 57.9 34.8 5-methyloxazolylperoxy 26.8 31.7 A 34.2 1000 K 2-methylfuranylperoxy 29.7 36.8 A 35.3 3-methylfuranylperoxy 39.8 34.8 42 40.9 2-methyloxazolylperoxy 28.6 68.8 A 35 4-methyloxazolylperoxy 31.4 26.5 56.4 32.4 5-methyloxazolylperoxy 25 30.6 A 31.5 aCyclization at this position did not result in a reasonable geometry.

Table 4.2. Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*) over the 298 to 2000 K range, for peroxy radical rearrangement pathways of the methylated furan and oxazole derivatives. Energies (kcal/mol) expressed relative to the starting peroxy radical (4.2). Table 4.2 continued on page 210.

214

Table 4.2, continued.

TS(4.2-4.3) TS(4.2-4.4a) TS(4.2-4.4b) TS(4.2-4.5) 1250 K 2-methylfuranylperoxy 27.6 35.4 a 32 3-methylfuranylperoxy 41.3 37.1 44.3 41.5 2-methyloxazolylperoxy 27.5 68.1 a 32.8 4-methyloxazolylperoxy 29.7 25.6 54.7 29.7 5-methyloxazolylperoxy 25.8 32.3 a 31.4 1500 K 2-methylfuranylperoxy 25.1 33.7 a 28.4 3-methylfuranylperoxy 42.8 39.5 46.6 42.1 2-methyloxazolylperoxy 26.6 67.3 a 30.6 4-methyloxazolylperoxy 27.9 24.6 52.8 26.9 5-methyloxazolylperoxy 20.8 28.1 a 25.5 1750 K 2-methylfuranylperoxy 22.5 33.5 a 24.6 3-methylfuranylperoxy 44.4 42 49 42.8 2-methyloxazolylperoxy 25.5 66.7 a 28.3 4-methyloxazolylperoxy 25.9 23.4 50.9 24 5-methyloxazolylperoxy 18.6 26.7 a 20.8 2000 K 2-methylfuranylperoxy 19.7 30 a 20.5 3-methylfuranylperoxy 46.1 44.5 51.5 43.6 2-methyloxazolylperoxy 24.4 65.8 a 26 4-methyloxazolylperoxy 24 22.2 48.8 20.9 5-methyloxazolylperoxy 16.2 25.2 a 19 aCyclization at this position did not result in a reasonable geometry.

As expected, the rearrangement pathways (4.2 → 4.3, 4.2 → 4.4, and 4.2 → 4.5) occurred with lower barriers and energies of reaction than did the dissociative pathways

(4.2 → 4.6 and 4.2 → 4.7). In particular, ipso cyclization (4.2 → 4.3) occurred most readily for the 2-substituted derivative, with an activation barrier of 30.2 kcal/mol and energy of reaction of 20.1 kcal/mol; for the 3-substituted derivative, ortho cyclization

(4.2 → 4.4a) occurred with the lowest barrier (28.9 kcal/mol) and energy of reaction

(13.8 kcal/mol). This is presumably due to the ability of these pathways to result in

215

allylic radical systems (Figure 4.5); an additional ortho cyclization (4.2 → 4.4b) was

possible for the derivative but was less favorable due to the lesser stability of the resultant

radical.

O O O O O O

O O O

O O H H O H

O O O

O O O

Figure 4.5. Allylic stabilization afforded via the most likely cyclization pathways of 2- methylfuranylperoxy radical (top) and 3-methylfuranylperoxy radical (bottom).

The barriers to ipso (4.2 → 4.3) and ortho cyclization (4.2 → 4.4) were seen to be

within 10 kcal/mol of one another, for both methylfuranylperoxy radicals; the activation

barrier for 1,3-H transfer (4.2 → 4.5) was also in this range. Interestingly, the products

of this transfer (as verified by displacement of the transition state geometry followed by

careful minimization to products) were the separated species 2- or 3-furanal and hydroxyl

(HO•) radical, such that the overall energy of reaction was highly favorable (~ −35 kcal/mol). Thus, at 298 K, cyclizations and H-atom transfer reactions are likely to dominate the chemistry of these species. The dissociative processes have considerably higher energies of reaction; however, at higher temperatures, they become more likely

216

due to the favorable entropy. (The relevant data are shown in Figures 4.16 and 4.17 and

will be discussed subsequently.)

The energies of the corresponding pathways of benzylperoxy radical are included

(Table 4.1) for comparison. Overall, the same pathways are competitive at 298 K. Of the

rearrangement pathways, ortho cyclization (4.2 → 4.4) has the lowest barrier (35.6 kcal/mol), followed by ipso cyclization (4.2 → 4.3, 37.2 kcal/mol) and 1,3-H transfer (4.2

→ 4.5, 38.9 kcal/mol). Since both ortho and ipso cyclization of benzylperoxy radical result in an allylic radical system, due to the lack of a heteroatom, the factor that dictates the most favorable pathway is ring size: a five-membered ring is more thermodynamically stable than a four-membered ring. This is also evident in the energies of reaction; ortho cyclization occurs with a free energy of reaction that is >10 kcal/mol more favorable than that of ipso cyclization. Notably, the functionalized furan derivatives all demonstrate cyclization barriers that are less endoergic than those of benzylperoxy radical. The barriers for the respective 1,3-H transfers are roughly equivalent; logically, the chemistry of the ring does not have a great effect on a process solely involving the side chain.

Relative to the pathways available to the non-alkylated heteroaromatic peroxy

radicals, these routes occurred with varying free energies of reaction. In particular, the

step that could be most directly compared was oxygen-atom loss (4.2 → 4.6), which

occurred with a free energy of reaction of only 3.4 kcal/mol in 2-furanylperoxy radical32

but nearly 50 kcal/mol in the methylated derivative. Likewise, ipso cyclization (4.2 →

4.3) occurred with ΔGo = 5.4 kcal/mol for 2-furanylperoxy radical, and ΔGo = 20.1

217

kcal/mol for the methylated derivative. It was thought that these differences in energy

could be attributed to the greater aromatic nature of the resultant non-methylated radicals.

Conversely, ortho cyclization (4.2 → 4.4) was more favorable for 2-methylfuranylperoxy

radical (ΔGo = 27.1 kcal/mol) than the non-alkylated compound (ΔGo = 36.4 kcal/mol),

presumably due to the respective size of the rings formed via these pathways. Thus, size

of the newly-cyclized ring and electronic delocalization in the newly-formed radicals

were seen to be the trends most important in dictating the preferred chemical pathways at

298 K for these peroxy radicals, with delocalization overshadowing the ring-size effect.

4.3.2. Oxazole derivatives. Structurally, the oxazole ring is similar to a furan ring, with

nitrogen in place of one of the carbons (Figure 4.1). Thus, the reactions of 2-

methyloxazolylperoxy radical and 5-methyloxazolylperoxy radical could both be compared to those of 2-methylfuranylperoxy radical, while the reactions of 4- methyloxazolylperoxy radical were best compared to 3-methylfuranylperoxy radical.

Moreover, this heteroaromatic ring allowed comparison of chemistry adjacent to an oxygen atom (5-derivative) to chemistry adjacent to a nitrogen atom (4-derivative), as well as to the influence of both O and N heteroatoms simultaneously (2-derivative).

Likely pathways are shown for the oxazolylperoxy derivatives (Figures 4.6-4.8).

Additionally, energies are compared to those of the relevant furan derivative (Table 4.3).

In general, the presence of the second heteroatom appeared to stabilize the relevant product, as the energies of reaction for the oxazolylperoxy derivatives were roughly 2 kcal/mol more endoergic than those of the furanylperoxy derivatives.

218

4.3

O O 4.4 O O N

21.2 CH2O N

O 7.7 33.6 72.7

4.1 4.5 O 4.2 O O 42.7 CHO +OH CH 2 +O2 CH2OO N N 1.8 N 0.0 -34.2

4.6 O +O(3P) CH2O 4.7 N O 50.9

+ O N O 54.3

Figure 4.6. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 2-methyloxazolylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

219

4.3 O 4.4a H O O O 17.5 N O 32.1 O 4.4b N

36.6 29.7 7.9 4.1 N 4.2 61.6 O O O O 39.8 N O 4.5 N OOH C H2C 2 1.1 +OH +O2 0.0 N OHC -35.2

O 4.6 +O(3P) 4.7 O O N OH2C + O 50.7

N 47.6

Figure 4.7. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 4-methyloxazolylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

220

4.3 O O O

21.1 N 4.4 O

31.3 O O N 24.7 34.6 4.1 H OOH C 4.2 H2C 2 O 4.5 O 39.7 OHC O OH +

N N 5.6 0.0 -37.7 +O2 N 4.6 OH2C O +O(3P)

N 53.0 4.7 O O O +

48.3 N

Figure 4.8. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 5-methyloxazolylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

221

2-methyloxazolylperoxy 2-methylfuranylperoxy 4.2 → 4.1 1.8 1.3 4.2 → 4.3 21.2 23.9 4.2 → 4.4 7.7 30.8 4.2 → 4.5 -34.2 -38.5 4.2 → 4.6 50.9 48.7 4.2 → 4.7 54.3 54.3

4-methyloxazolylperoxy 3-methylfuranylperoxy 4.2 → 4.1 1.1 3.0 4.2 → 4.3 32.1 33.6 4.2 → 4a 17.5 11.6 4.2 → 4b 7.9 31.6 4.2 → 4.5 -35.2 -36.2 4.2 → 4.6 50.7 48.3 4.2 → 4.7 47.6 52.4

5-methyloxazolylperoxy 2-methylfuranylperoxy 4.2 → 4.1 5.6 1.3 4.2 → 4.3 21.1 23.9 4.2 → 4.4a 24.7 30.8 4.2 → 4.5 -37.7 -38.5 4.2 → 4.6 53.0 48.7 4.2 → 4.7 48.0 54.3

Table 4.3. Exploration of multiple heteroatom effect through comparison of the methyloxazolylperoxy derivatives to the corresponding methylfuranylperoxy derivative. Free energies of reaction (kcal/mol) determined at the B3LYP/6-311+G**//B3LYP/6- 31G* level of theory at 298 K.

Overall, the qualitative trends were repeated between the two sets of derivatives. One exception occurred, in the case of oxygen loss versus side chain loss. Atomic oxygen loss (4.2 → 4.6) was more favorable for the furan derivatives, while dioxirane loss (4.2

→ 4.7) was favored for the oxazole derivatives. This indicated that the ring-centered oxazolyl radicals were more stable than the furanyl radicals; this was attributed to the

222

effect of the nitrogen atom: as formation of the heteroaromatic radical (4.2 → 4.7) was more heavily influenced by the identity of the atoms in the ring than other pathways, it is unsurprising that its results would be more greatly affected.

4-Methyloxazolylperoxy radical is substituted adjacent to the ring’s nitrogen

atom; 5-methyloxazolylperoxy radical is substituted adjacent to the oxygen atom. Ipso

cyclization (4.2 → 4.3) of the 4-substituted derivative occurred with a higher activation

barrier and higher free energy of reaction than did ipso cyclization of the 5-substituted

derivative. Conversely, 4-methyloxazolylperoxy radical saw a lower barrier and energy

of reaction for ortho cyclization (4.2 → 4.4a) than did 5-methyloxazolylperoxy radical.

These preferences are attributed to the same causes seen in the monoheterocyclic

derivatives. (This was validated by the case of ortho cyclization for the two species, as a

five-membered ring was formed in both cases, so the major difference between the 4.2 →

4.4a products was the degree of allylic resonance.) For 4-methyloxazolylperoxy radical,

Pathway 4.2 → 4.4a resulted in more allylic stabilization than did pathway 4.2 → 4.3,

while for 5-methyloxazolylperoxy radical, the reverse was true.

Due to the nature of the oxazole ring, ortho cyclization at nitrogen is a possibility

for both 2-methyloxazolylperoxy radical and 4-methyloxazolylperoxy radical. In our

previous work on the azabenzenes,33 we postulated that this might be a pathway with implications for NOx chemistry but that the barrier to such a reaction was prohibitive at

298 K. This was repeated here, with activation barriers of 72.7 kcal/mol for the 2-

substituted compound and 61.6 kcal/mol for the 4-substituted derivative, over 30

kcal/mol above the barriers of the likely cyclizations.

223

4.3.3. Pyrrole derivatives

The chemistry of the pyrrole derivatives is perhaps most relevant to our previous work on the azabenzenes; additionally, the nitrogen atom allows for additional pathways of interest (Figures 4.9 and 4.10).

4.3 H 4.4a N O H N O

25.1 O H O 4.4b 29.1 OOH N 29.1 35.5

10.7 H 4.1 H 4.2 22.4 N N

CH +O CH OO H 4.5 2 2 2 38.8 N CHO 3.5 0.0 +OH

-42.7

H 4.6 N +O(3P) CH2O H 4.7 N 47.1 + O O 52.2

Figure 4.9. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 2-methylpyrrolylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

224

H N 4.4a H N 4.3 H

O 21.1 35.4 O O H O N

4.4b 30.2 36.9 H O 33.0 36.8 O 4.1 4.2 H H N N 38.8 H 4.5 +O2 N +OH 8.0 0.0 CH 2OO CH2 -39.3 CHO

H N 4.6 +O(3P) 4.7 H N O CH2O 49.3 + O

51.5

Figure 4.10. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 3-methylpyrrolylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

Overall, similar trends were seen. Ipso cyclization (4.4 → 4.3) was roughly 4 kcal/mol more favorable than ortho cyclization at carbon (4.4 → 4a). 1,3-Hydrogen-atom transfer

225

(4.4 → 4.5), followed by dissociation to products, occurred with a barrier competitive

with that of ipso cyclization. Additionally, the structure of this derivative allowed a

second H-atom transfer, via an ortho cyclization at nitrogen (4.4 → 4.4b), which saw the

lowest barrier overall, attributed to the stability of the resultant N-centered radical.

Since oxidation of this resulting N-centered radical could seemingly result in the

formation of NOx species, this pathway was subjected to further exploration (Figure

4.11). It was seen that this radical could dissociate to formaldehyde, hydroxyl radical,

and a high-energy, diradical “pyrrynyl” intermediate, at a high energetic cost (60.1

kcal/mol); another pathway entailed oxidation at the incipient radical center, with a free

energy of reaction of 11.2 kcal/mol. This latter pathway could then proceed through

several of the same peroxy radical pathways discussed here, such as hydrogen-atom

transfer, rearrangement, or dissociation to various products. The fact that oxidation at the nitrogen atom was itself endoergic was notable, as oxidation at a carbon-centered radical was exoergic for all the species explored in this work. It is borne out by another study completed in our group, which demonstrated that reactive species centered on nitrogen are generally more stable that those centered on carbon;34 thus, it follows that the nitrogen-centered radical would be less reactive than a carbon-centered radical.

226

N O + +OH N H H CH2OOH 60.0

O 0.0 2 O O 34.1 rearrangement N pathways

CH2OOH

11.2

Figure 4.11. Potential steps available to the N-centered radical formed via ortho cyclization / hydrogen-atom transfer in 2-methylpyrrolylperoxy radical. Free energies of reaction and free energy of activation for oxygen addition (B3LYP/6- 311+G**//B3LYP/6-31G*, kcal/mol, 298 K) are relative to the starting radical.

3-Methylpyrrolylperoxy radical favored cyclization at the ortho position (4.2 →

4.4a) by roughly 14 kcal/mol over cyclization at the ipso (4.2 → 4.3) position. Ipso

cyclization and the 1,3-H transfer follow with comparable barrier heights (the N−H

abstraction pathway for this derivative presumably occurs with a high degree of strain in

the transition state, and the relevant geometry was not found). Overall, these trends

matched those of the other 3-substituted heterocyclic peroxy radicals.

4.3.4. Thiophene derivatives. The energetic preferences of 2-methylthiophenylperoxy radical and 3-methylthiophenylperoxy radical were explored at 298 K (Figures 4.12 and

4.13). Overall, energetic preferences track those of the other monocyclic heteroaromatic

227

rings; ipso and ortho cyclization dominate the energetic preferences of the 2- and 3- substituted methylthiophenylperoxy radicals, respectively.

4.3 4.4a S O S O

22.9 O H O 4.4b O 24.2 O 33.9 S 31.7

48.6 14.1 4.1 4.2 S S

CH2 +O2 CH OO 2 38.3 S 4.5 4.2 0.0 CHO +OH

-39.5

S 4.6 +O(3P) CH2O 4.7 S 47.6 + O O 48.0

Figure 4.12. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 2-methylthiophenylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

228

S 4.4a S 4.3 H

O 15.9 O 32.8 O O S

4.4b 30.0 35.9 H O 27.4 33.5 O 4.1 4.2 S S 38.9 4.5 +O2 S +OH 5.6 0.0 CH OO CH 2 2 -39.1 CHO

S 4.6 +O(3P) 4.7 S 47.8 O CH2O + O

45.3

Figure 4.13. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) of the rearrangement pathways for 3-methylthiophenylperoxy radical (4.2), shown relative to the starting peroxy radical. Activation barriers are shown for the likely pathways.

2-Methylthiophenylperoxy radical still showed a preference for ipso cyclization and the allylic stabilization this pathway entailed, but the preference of one cyclization

229

route over another was less pronounced for these species than the other heterocyclic

derivatives. Ortho cyclization (4.2 → 4.4a) was only 1.3 kcal/mol more endoergic, in

terms of the free energies of reaction. A second ortho substitution (4.2 → 4.4b) was

available to 2-methylthiophenylperoxy radical, which resulted in the formation of an S−O

radical, which could have implications in SOx formation. This process occurred with an

activation barrier of 33.5 kcal/mol and a free energy of reaction of 27.4 kcal/mol, so is

unlikely to contribute to SOx formation at 298 K. 3-Methylthiophenylperoxy radical

demonstrated energetics comparable to all other 3-substituted derivatives, favoring ortho

cyclization adjacent to the heteroatom (4.2 → 4.4a).

4.3.5. Overall low-temperature trends

Overall, the derivatives of the monocyclic heteroatomic derivatives (furan,

pyrrole, and thiophene) compared well in their chemistry and preferred pathways; the presence of a second heteroatom, in oxazole affected chemistry only at the 2-position,

where the effects of the two heteroatoms would be most pronounced.

Comparing the results among the methyl-substituted five-membered

heteroaromatics, some overall trends can be discerned. The driving force behind the peroxy radical rearrangements at 298 K is the possibility of forming an allylic radical system upon cyclization. The presence of a heteroatom disrupts the allylic system; this was postulated for some related reactions of the azabenzenes but is more evident in these systems. Thus, the 2-substituted derivatives favor ipso cyclization, while the 3- substituted derivatives prefer ortho cyclization. This allylic stabilization supersedes any

230

other factors which might preference a given stabilization; whereas with the azabenzenes,

ortho cyclization was consistently favored due to its yielding a five-membered ring, for

these smaller heteroaromatics, the energetic benefit of the pathway that best maintained

the allylic system overcame any detrimental effects of forming a strained ring.

Regardless of substitution, the H-atom transfer route had a comparable barrier, roughly

38−39 kcal/mol. Because this reaction takes place on the alkyl chain, the radical center

cannot communicate with the aromatic system and thus the energetics are consistent.

These results were also seen with the methylazabenzylperoxy radicals, for the same

reason.

Unlike the unsubstituted heteroaromatics’ peroxy radicals, these species did not

lose atomic oxygen or their side chains without significant energetic cost. This can be

attributed to the lack of aromaticity in the end product (Figure 4.14). Whereas 2-

furanyloxy radical (A) achieves a degree of π delocalization upon loss of an oxygen atom

from 2-furanylperoxy radical, such delocalization is not possible for the 2-

methylfuranyloxy radical (B). Thus, the energy of reaction simply reflected the breaking of an O−O bond, generally worth ~50 kcal/mol, and was comparable in the case of all derivatives explored. Similarly, side-chain loss for 2-furanylperoxy radical is the

dissociation to the aromatic radical and molecular oxygen and occurs readily; for 2-

methylfuranylperoxy radical, the side chain would be lost either as a strained three-

membered ring (dioxirane) or a high-energy diradical, neither of which would be particularly stable.

231

O

O

O O

O CH2O

B

O

O

A

Figure 4.14. Comparison of 2-furanylperoxy radical (A) to 2-methylfuranylperoxy radical (B). Additional resonance stabilization possible for A due to oxygen’s increased proximity to the heteroaromatic.

The five-membered heteroaromatics afforded the opportunity to observe the

effects of heteroatom substitution, which was not available in the azabenzenes. For

instance, each 2-derivative favored ipso cyclization (4.2 → 4.3) over ortho cyclization, to

best maintain its allylic system. However, the degree of preference varied between these

species: 2-methylfuranylperoxy radical favored ipso cyclization by 6.9 kcal/mol; 2- methylpyrrolylperoxy radical, by 4.0 kcal/mol; and 2-methylthiophenylperoxy radical, by

1.3 kcal/mol. Similarly, in the case of the 3-derivatives, where ortho cyclization (4.2 →

232

4.4a) was favored over ipso cyclization, 3-methylfuranylperoxy radical favored ortho

cyclization by 22.0 kcal/mol; 3-methylpyrrolylperoxy radical, by 14.3 kcal/mol; and 3-

methylthiophenylperoxy radical, by 16.9 kcal/mol.

4.3.6. Heteroatom effects: comparisons to benzylperoxy radical

Relative to the hydrocarbon equivalent benzylperoxy radical, a few interesting

trends were discerned from the data of the five-membered heteroaromatic rings. While

the qualitative trends were comparable regardless of ring size and/or the presence of a

heteroatom, the quantitative energies varied with pathway. As mentioned previously,

ipso cyclization (4.2 → 4.3) for benzylperoxy radical did not yield a particularly stable

product; however, in the 2-derivatives, this pathway yielded an allylic derivative that

ortho cyclization (4.2 → 4.4) could not provide. Thus, the energies of ipso cyclization

for all of the peroxy radicals derived from the 2-substituted heteroaromatics are notably

lower than that of benzylperoxy radical (Table 4.4). Conversely, ortho cyclization (4.2

→ 4.4a) was generally more favorable for benzylperoxy radical than for the 3-substituted

heteroaromatic peroxy radicals, with the lone exception being 3-methylfuranylperoxy

radical. For the relevant derivatives (2-methylpyrrolylperoxy, 4-methyloxazolylperoxy,

and 4-methyloxazolylperoxy radicals), ortho cyclization at a heteroatom (4.2 → 4.4b)

occurred with a lower energy of reaction, but the relevant barriers to reaction were

considerably higher. As might be expected, the remaining pathways (4.2 → 4.1, 4.2 →

4.5, 4.2 → 4.6, and 4.2 → 4.7) were similar in their energetics regardless of whether the starting peroxy radical resulted from benzyl radical or a heterocyclic radical: these

233

pathways did not involve the chemistry of the central ring to any great extent. Thus, the

presence of a heteroatom did affect the likely reaction pathways to some extent,

particularly with respect to rearrangement routes; moreover, these heteroatoms influenced

the range of products that could be generated via internal reaction.

Enthalpies of Reaction 4.2 → 4.1 4.2 → 4.3 4.2 → 4.4a 4.2 → 4.4b 4.2 → 4.5 4.2 → 4.6 4.2 → 4.7 Furan 11.5 18.4 24.6 -29.3 57.6 64 Pyrrole 13.8 24.4 27.9 11.8 -32.9 56.4 64.6 Thiophene 14.4 21.8 22.3 14.4 -29.9 57.1 59.7

Free Energies of Reaction 4.2 → 4.1 4.2 → 4.3 4.2 → 4.4a 4.2 → 4.4b 4.2 → 4.5 4.2 → 4.6 4.2 → 4.7 Furan 3.5 25.1 29.1 -42.7 47.1 52.2 Pyrrole 1.3 20.1 27.1 10.7 -38.5 48.7 52.7 Thiophene 4.2 22.9 24.2 14.1 -39.5 47.6 48

Enthalpies of Activation TS(4.2-4.3) TS(4.2-4.4a) TS(4.2-4.4b) TS(4.2-4.5) Furan 28.5 32.8 -- 37.8 Pyrrole 28.0 33.8 21.6 38.6 Thiophene 30.3 31.9 46.7 37.7

Free Energies of Activation TS(4.2-4.3) TS(4.2-4.4a) TS(4.2-4.4b) TS(4.2-4.5) Furan 29.1 35.5 -- 38.8 Pyrrole 30.2 35.2 22.4 38.8 Thiophene 31.7 33.9 48.6 38.3

Table 4.4. Comparing free energies of activation and reaction for 2-substituted methyl peroxy radical derivatives of furan, pyrrole, and thiophene, relative to starting peroxy radicals (4.2). All quantities in kcal/mol (B3LYP/6-311+G**//B3LYP/6-31G*) at 298 K.

234

4.3.7. Ring size comparison: Heteroaromatics

An additional comparison of interest involves the chemistry of these peroxy

radicals relative to that of the azabenzenes, on which we previously reported (Table 4.5).

Generally, the rearrangement pathways occurred with lower energies for the five-

membered heteroaromatics. This was especially notable in the case of the most favorable rearrangement for each species. Regardless of substitution, ipso cyclization (4.2 → 4.3)

was more favorable for the five-membered rings than the azabenzenes; in all but a few

cases, the same was true for ortho cyclization (4.2 → 4.4). The exception to this latter

statement was sometimes seen when ortho cyclization did not result in an allylic radical

system. Furthermore, the key difference between the chemistries of these two ring

systems was seen in the preferences between these pathways: the azabenzylperoxy

radicals always preferred ortho cyclization, while the five-membered heterocyclic peroxy

radicals preferred the cyclization that would yield an allylic radical system.

235

Parent Diazabenzene Substitution 4.1 4.3 4.4a 4.4b 4.5 4.6 4.7 Pyridine 2 6.5 35.7 23.2 -4.1 -39.4 53.0 38.7 3 5.6 34.8 23.8 26.8 -38.3 49.3 43.2 4 5.6 37.5 24.5 24.5b -37.4 53.8 41.6 Pyridazine 3 6.0 36.5 22.2 -9.1 -32.4 50.5 39.5 4 4.1 35.1 24.7 24.0 -36.8 50.5 39.8 Pyrimidine 2 6.3 37.5 0.2a 0.2b -33.6 52.5 40.4 4 5.8 39.8 25.3 1.1 -31.5 50.9 38.4 5 8.2 33.9 26.2 26.2b -35.2 51.6 45.6 Pyrazine 2 4.4 33.3 21.3 -6.1 -34.1 50.4 37.5 a For 2-methylpyrimidinylperoxy radical, both 4a and 4b refer to ortho cyclization at nitrogen. bOrtho pathways equivalent via symmetry.

Table 4.5. Free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reaction pathways available to the peroxy radicals of the azabenzenes. Where applicable, 4.4a refers to cyclization at carbon and 4.4b refers to ortho cyclization at nitrogen. All reaction pathways relative to the peroxy radical (4.2) of the relevant parent compound.

The energies of hydrogen-atom transfer (4.2 → 4.5) corresponded well regardless

of ring size, as did the energies of atomic oxygen loss (4.2 → 4.6) and molecular oxygen

addition (4.2 → 4.1). Loss of dioxirane (4.2 → 4.7) was notably less endoergic for the

methylazabenzylperoxy derivatives than for the five-membered heterocyclic derivatives,

presumably due to the increased aromatic nature of the resulting radicals for the six-

membered rings.

The most notable difference in reactivity between the two sets of heteroaromatics involved the increased likelihood of certain cyclizations. This could have implications on

further decomposition products. For instance, it can be hypothesized that the ipso

cyclization products could readily dissociate into formaldehyde and the corresponding

oxyradical (Figure 4.15), while the ortho cyclizations would likely yield different sorts of

decomposition products.

236

X O X X

O OOCH2 O 4.3

+CH2O

X X X 4.5 CHO CH2OO CHOOH

+OH 4.6

X 4.7 3 CH2O +O( P)

X O O

Figure 4.15. Peroxy radical pathways resulting in aryl or aryloxy radicals either directly (4.2 → 4.6, 4.2 → 4.7) or indirectly (4.2 → 4.3), along with pathway resulting in carbonyl-functionalized compound (4.2 → 4.5). These subsequent steps available could build on extant combustion pathways known for likewise functionalized compounds.

This could ultimately have implications in terms of predicting coal chemistry; depending on the size and prevalence of the ring units in a sample, different products would be more likely than others. Overall, likely next steps for these species would largely overlap with previous work in our group, as several pathways result in either aryl or aryloxy radicals, which we have studied.

4.3.8. Temperature effects

Free energies of reaction for these decomposition steps vary greatly with

temperature. Dissociative steps tend to see a large entropic benefit at higher temperatures

that rearrangements would not experience; thus, different pathways are likely to dominate

the chemistry of these peroxy radicals at different temperatures. The free energies at

237

temperatures ranging up to 2000 K were calculated for each of the methylated

heterocyclic peroxy radicals (Table 4.1). These data are summarized graphically for 2-

methylfuranylperoxy radical (Figure 4.16). As can be seen, the purely rearrangement

pathways (4.2 → 4.3 and 4.2 → 4.4) varied only slightly over the temperature range and did not vary in their relative relationship; ipso cyclization (4.2 → 4.3) was consistently more favorable. The purely dissociative pathways (4.2 → 4.1, 4.2 → 4.6, and 4.2 → 4.7) saw a much greater temperature effect, dropping precipitously as the temperature increased. Generally, dissociation back to reactants (4.2 → 4.1) quickly became exoergic

(at T > 500 K), while loss of oxygen atom and loss of dioxirane became exoergic in the high-temperature combustion range (T ~1250-1750 K). Finally, the 1,3-H transfer (4.2

→ 4.5) was highly exoergic regardless of temperature.

238

Figure 4.16. Variation in ΔGrxn (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol) with temperature for the pathways available to 2-methylfuranylperoxy radical. 4.2 → 4.1 denoted by solid diamond; 4.2 → 4.3 denoted by solid square; 4.2 → 4.4 denoted by solid triangle; 4.2 → 4.5 denoted by x; 4.2 → 4.6 denoted by asterisk; 4.2 → 4.7 denoted by open circle.

Barrier heights for the rearrangement pathways (4.2 → 4.3, 4.2 → 4.4, and 4.2 →

4.5) were also explored as a function of temperature (Table 4.2), for several of the

compounds of interest. Again using 2-methylfuranylperoxy radical as a representative

compound (Figure 4.17), it can be seen that all barrier heights decreased over this

temperature range. The largest decrease occurred for 1,3-H transfer (4.2 → 4.5), given the dissociative nature of this pathway. These results were mirrored in the overall activation energies of the furanyl and oxazolyl derivatives. In particular, for both 3- methylfuranylperoxy radical and 4-methyloxazolylperoxy radical, 1,3-H transfer (4.2 →

4.5) saw such an entropic benefit that it demonstrated the lowest barrier at temperatures higher than 1500 K.

239

Figure 4.17. Variation in ΔGactivation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol) with temperature for the most energetically favorable pathways available to 2- methylfuranylperoxy radical. TS(4.2-4.3) denoted by solid diamond; TS(4.2-4.4) denoted by solid square; TS(4.2-4.5) denoted by open triangle.

The rearrangement pathways demonstrated the lowest activation barriers far more

consistently in this work than in our previous exploration of the methylazabenzylperoxy

radicals, which saw the 1,3-H transfer pathway become more favorable at T > 1250 K.

Both the increased stability of the cyclization products as well as a slightly lesser

energetic benefit of dissociation contributed to this trend. Nonetheless, given the “dual

nature” of this pathway, it appeared to be a factor at both low and high combustion

temperatures: as a hydrogen-atom transfer, it participated in low-temperature chemistry,

but given the quick dissociation to products, it saw a preferential benefit with increasing

240

temperature. Thus, it is likely to be a contributing pathway regardless of temperature range, and the resultant products could be significant products of heteroaromatic combustion.

4.3.9. Ethyl derivatives

The effect of increasing alkyl substitution was explored via examination of the relative kinetics and thermodynamics for both the methyl and ethylfuranylperoxy radicals

(Table 4.6). The overall thermodynamics for all the ethyl-substituted heteroaromatics were also of interest (Table 4.7). It can be seen that the ethyl derivatives demonstrated lower activation barriers and reaction energies in most potential pathways. Generally, this could be rationalized in terms of destabilization of the initial radicals or stabilization of the relevant transition state, relative to the methyl derivatives. Given that radical stabilization increases with the amount of alkyl substitution, it seems more likely that the additional alkyl group present in the ethyl derivatives stabilizes both the products and the transition states leading to them. This was in keeping with our initial studies of the bond dissociation enthalpies of methyl- and ethyl-substituted heteroaromatics, which suggested that increasing alkyl substitution led to increased reactivity. Heteroatom effect was explored, but generally, quantitative effects of the additional alkyl group were comparable. Moreover, the quantitative effect was insignificant (<2 kcal/mol) for all rearrangements, and small (<5 kcal/mol) across all reaction pathways. Finally, qualitative trends were duplicated between the methyl and ethyl derivatives. Thus, the

241

smaller methyl derivatives can provide a useful model for the chemistry of other alkylated heteroaromatics.

2-Substituted 3-Substituted Methyl Ethyl Methyl Ethyl 4.2 → 4.1 1.3 0.9 3.0 4.9 TS(4.2-4.3) 30.2 27.0 36.2 33.7 4.2 → 4.3 20.1 17.7 33.6 31.0 TS(4.2-4.4a) 35.2 33.6 28.9 26.0 4.2 → 4.4a 27.1 25.7 13.8 14.2 TS_4b a a 36.3 34.4 4.2 → 4.4b a a 31.6 29.9 TS_5 38.8 37.2 39.2 39.9 4.2 → 4.5 -38.5 -43.2 −36.2 −40.6 4.2 → 4.6 48.7 50.0 48.3 46.4 4.2 → 4.7 52.7 48.9 52.4 46.2 aCyclization did not occur at this position.

Table 4.6. Variation in free energies (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298 K) with increasing alkyl substitution (kcal/mol) at 298 K, for the methyl and ethyl- substituted 2- and 3-alkylfuranylperoxy radicals.

242

2-Substituted 3-Substituted Pyrrole Derivatives Methyl Ethyl Methyl Ethyl 4.2 → 4.1 3.5 3.2 8.0 7.5 4.2 → 4.3 25.1 22.2 35.4 33.2 4.2 → 4.4a 29.1 28.6 21.1 20.0 4.2 → 4.4b 10.7 9.4 33.0 31.5 4.2 → 4.5 -42.7 -47.9 -39.3 -43.4 4.2 → 4.6 47.1 47.7 49.3 50.3 4.2 → 4.7 52.2 45.3 51.5 44.9

2-Substituted 3-Substituted Thiophene Derivatives Methyl Ethyl Methyl Ethyl 4.2 → 4.1 4.2 1.0 5.6 5.2 4.2 → 4.3 22.9 21.5 32.8 15.1 4.2 → 4.4a 24.2 24.8 15.9 31.1 4.2 → 4.4b 14.1 12.8 27.4 26.6 4.2 → 4.5 -39.5 -43.0 -39.1 -39.8 4.2 → 4.6 47.6 49.6 47.8 49.8 4.2 → 4.7 48.0 42.4 45.3 39.5

2-Substituted 4-Substituted 5-Substituted Oxazole Derivatives Methyl Ethyl Methyl Ethyl Methyl Ethyl 4.2 → 4.1 1.8 1.4 1.1 5.3 5.6 0.3 4.2 → 4.3 21.2 18.7 32.1 30.3 21.1 18.3 4.2 → 4.4a 7.7 8.3 17.5 24.0 24.7 21.7 4.2 → 4.4b a a 7.9 9.0 a a 4.2 → 4.5 -34.2 -38.7 -35.2 -38.7 -37.7 -43.2 4.2 → 4.6 50.9 49.0 50.7 48.8 53.0 49.2 4.2 → 4.7 54.3 45.2 47.6 44.8 48.3 46.6 aCyclization did not occur at this position.

Table 4.7. Variations in free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298 K), with varying substitution for the alkylfuranylperoxy, alkylthiophenylperoxy, and alkyloxazolylperoxy radicals.

4.3.10. Oscillator vs. rotor

The effects of generating the thermal corrections to the enthalpy and free energy via a rigid rotor approximation rather than the harmonic oscillator approximation were examined (Table 4.8).

243

Parent Methyl 4.1 4.3 4.4a 4.4b 4.5 4.6 4.7 Heteroaromatic Substitution Furan 2 2.6 21.4 28.4 a -37.2 49.6 54.0 3 3.9 35.0 15.2 33.1 -37.0 49.2 51.7 Oxazole 2 3.0 22.4 9.0 a -32.9 62.9 52.2 4 0.9 32.6 18.1 8.4 -34.6 56.7 51.3 5 7.1 22.7 26.3 a -36.0 57.6 54.7 Pyrrole 2 3.0 25.5 29.5 11.2 -42.3 42.1 52.6 3 9.4 35.4 21.1 33.0 -39.3 50.3 51.4 Thiophene 2 4.8 24.0 25.3 15.2 -38.6 47.4 49.1 3 6.7 34.0 17.1 28.6 -37.8 48.3 46.5 aCyclization did not occur at this position

Table 4.8. Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298 K), using the rigid rotor approximation in place of oscillator approximation.

It was shown that effects on the quantitative energies were minimal (<2 kcal/mol) and the qualitative energetic preferences of the species remained the same. As expected, a greater effect was seen with the free energies; information on the enthalpies of reaction and activation barriers at 298 K is included in the Supporting Information. Overall, it was shown that using the harmonic oscillator approximation did not introduce significant error to our understanding of the preferential reaction pathways.

4.4. Conclusions

Using density functional theory (B3LYP/6-311+G** //B3LYP/6-31G*), we examined the unimolecular decomposition steps available to the peroxy radicals of the alkylated, five-membered heteroaromatics. We saw that likely pathways at 298 K involved: (a) reversion to reactants; (b) rearrangement to form an allylic radical system;

244

or (c) 1,3-H transfer on the alkyl side chain, followed by scission to hydroxyl radical and

an aldehydic product. Presumably, subsequent steps with atmospheric implications

would involve: (a) decomposition of the unoxygenated heterocyclic radical; (b) further

decomposition or rearrangement of the newly-cyclized derivatives; or (c) reactions

common to hydroxyl radical and carbonyl-functionalized atmospheric species.

Temperature effects were explored, examining free energies of reaction from 298

to 2000 K. The energetics of the rearrangement pathways did not vary greatly over this

3 range; the dissociative pathways of atomic oxygen (O( P) loss, molecular oxygen (O2) loss, and dioxirane (c-CH2OO) side-chain loss saw a substantial entropic benefit and

became exoergic at higher temperatures. Internal H-atom transfer followed by scission to

a functionalized aldehyde and hydroxyl radical demonstrated aspects of both low-

temperature and high-temperature chemistry, with a relatively low barrier at 298 K but a

substantial entropic benefit at higher temperatures. Similarly, the free energies of

activation were explored for a select number of the rearrangement pathways; these barrier

heights dropped over the temperature range for all pathways but saw the largest effect for

the hydrogen-atom transfer route. Thus, as expected, the decomposition pathways would

likely play a more substantial role in high-temperature combustion, with these peroxy

radicals, themselves reactive oxygen species (ROS), also generating atomic oxygen and

hydroxyl radical.

This study allowed us to compare and contrast our results with other relevant

work in our group. Relative to the non-alkylated heteroatomic arylperoxy radicals, we

saw that our target species demonstrated a greater affinity for intramolecular reactions,

245

due to the increased length of their side chains. Dissociative reactions were less favorable for the alkylated derivatives, attributed to the lesser aromatic character present in the relevant pathways. With respect to the alkylated diazabenzylperoxy radicals, it was seen that, while cyclizations were favored pathways at 298 K regardless of parent ring size, the species with five-membered parent rings cyclized in such a manner as to

generate an allylic radical system (cyclizing at a position either ipso or ortho to the

substituted side chain), while the species with six-membered parent rings preferred to form five-membered rings via ortho cyclization, due to their increased stability over four- membered rings (which would result from ipso cyclization).

Given the variety of heteroatoms present in these species, it was possible to

observe their effects on the reactions of interest. The energetics of the heteroaromatics

were compared among themselves and to benzylperoxy radical. Qualitative trends in

preferred pathways were duplicated regardless of heteroatom, and quantitative energies

likewise were comparable between the alkylated heterocyclic peroxy radicals and

benzylperoxy radicals. Oxygen, nitrogen, and sulfur atoms did affect the chemistry of

their relevant heteroaromatics to different extents. Similarly, the effect of multiple

heteroatoms (as in oxazole) did affect the reaction energetics to an extent, even while

comparable qualitative trends were seen between all of the heteroaromatics. Notably,

the pyrrole, oxazole, and thiophene derivatives reacted to form various species wherein

their heteroatoms were oxidized, and these effects were unique to each heteroaromatic.

Thus, hydrocarbon species can be used to estimate the energetics of these species, but

246

cannot always be used to fully approximate their chemistry, due to the different products that are generated due to the heteroatoms present.

The effect of increasing alkyl substitution was examined through a comparison of

the peroxy radicals of the methyl and ethyl derivatives of the five-membered

heteroaromatics. It was seen that increasing alkyl substitution increased the reactivity of

these species, leading to lower activation barriers and lower free energies of reaction.

This corresponded well to our earlier findings with respect to these compounds. Finally,

the validity of the harmonic oscillator approximation, used in our calculations to generate thermal corrections to the enthalpy and entropy, was studied via comparison of our original results to the results obtained using the rigid rotor approximation. The resulting

comparable values legitimized our original calculations.

In summary, we have used the B3LYP method to study several aspects of the combustion pathways of the alkylated heteroaromatics. These results represent an important step in that study, providing details about the oxidative decomposition of the alkylated heteroaromatics- in particular, the peroxy radical pathways that have implications for these species across a wide temperature range. We have shown that the alkylated five-membered heteroaromatics demonstrate certain unique tendencies in their reactivity. Given the differences in reactivity between the five- and six-membered heteroaromatics, as well as between the alkylated and non-alkylated heteroaromatics, it seems likely that the chemical behavior of coal varies somewhat depending on the amounts and nature of the cyclic subunits present in the structure of a given sample.

247

References for Chapter 4

1 Smith, K. L.; Smoot, L. D.; Fletcher, T. H.; Pugmire, R. J. The Structure and Reaction

Processes of Coal; Plenum Press: New York, 1994.

2 Mackie, J.; Colket, M.; Nelson, P. J. Phys. Chem. 1990, 94, 4099-4106.

3 Kiefer, J.; Zhang, Q.; Kern, R.; Yao, J.; Jursic, B. J. Phys. Chem. B 1997, 101, 7061.

4 (a) Molto, J.; Font, R.; Conesa, J.A. Energ. Fuels 2006, 20, 1951-1958. (b) Ciccioli, P.;

Brancaleoni, E.; Frattoni, M.; Cecinato, A.; Pinciarelli, L. Anal. Lett. 2001, 34, 937-955.

(c) Sidhu, S.; Graham, J.; Stribech, R. Chemosphere 2001, 42, 681-690.

5 (a) McDonald, J. D.; Zielinska, B.; Fujita, E. M.; Sagebiel, J. C.; Chow, J. C.; Watson,

J. G. Env. Sci. Technol. 2000, 34, 2080-2091. (b) Ozawa, T.; Ishiguro, Y.; Toyoda, K.;

Nishimura, M.; Sasahara, T.; Doi, T. Chem. Sensors 2004, 20, 56-57.

6 (a) Levendis, Y. A.; Atal, A.; Carlson, J. B. Combust. Sci. Technol. 1998, 134, 407-431.

(b) Lemieux, P. M.; Ryan, J. V. Air and Waste 1993, 43, 1106-1115.

7 Baker, R. R.; Bishop, L. J. J. Anal. Appl. Pyrol. 2005, 74, 145-170.

8 (a) Xie, H.; Wang, J.; Zhang, S.; Ding, Y.; Sun, C. J. Chem. Phys. 2006, 125, 124317-

124317-10. (b) van der Laarse, J. D.; van Leuven, H. C. E. Anal. Chim. Acta 1966, 34,

370-371.

9 Bruinsma, O. S.; Tromp, P. J. J.; DeSauvage Nolting, H. J. J.; Mouljin, J. A. Fuel 1988,

67, 334-340.

10 Cullis, C. F.; Norris, A. C. Carbon 1972, 10, 525-537.

11 Braslavsky, S.; Heicklin, J. Chem. Rev. 1977, 77, 473-511.

12 Klein, E.; Lukes, V.; Cibulkova, Z.; Polovkova, J. J. Mol. Struct. 2006, 758, 149-159.

248

13 (a) Grela, M. A.; Amorebieta, V. T.; Colussi, A. J. J. Phys. Chem. 1985, 89, 38-41. (b)

Fulle, D.; Dib, A.; Kiefer, J. H.; Zhang, Q.; Yao, J.; Kern, R. D. J. Phys. Chem. A 1998,

102, 7480-7486. (c) Sendt, K.; Bacskay, G.B.; Mackie, J. C. J. Phys. Chem. 2000, 104,

1861-1875. (d) Organ, P. P.; Mackie, J. C. Chem. Soc. Faraday Trans. 1991, 87, 815-

823.

14 (a) Lifshitz, A.; Tamburu, C.; Suslensky, A. J. Phys. Chem. 1989, 93, 5802-5808. (b)

Bacskay, G. B.; Martoprawiro, M.; Mackie, J. C. Chem. Phys. Lett. 1998, 290, 391-398.

(c) Mackie, J. C.; Colket, M. B.; Nelson, P. F.; Esler, M. Int. J. Chem. Kinet. 1991, 23,

733-760.

15 (a) Parr, R. G.; Yang, W. Density Functional Theory in Atoms and Molecules; Oxford

University Press: New York, 1989. (b) Labanowski, J. W.; Andzelm, J. Density

Functional Methods in Chemistry; Springer: New York, 1991.

16 Barckholtz, C.; Barckholtz, T.; Hadad, C. J. Am. Chem. Soc. 1999, 121, 491-500.

17 Fadden, M. and Hadad, C. J. Phys. Chem. A 2000, 104, 6088-6094; Fadden, M. and

Hadad, C. J. Phys. Chem. A 2000, 104, 6324-6331.

18 Hayes, C. J.; Hadad, C. M. J. Phys. Chem. A, submitted.

19 Hayes, C. J.; Hadad, C. M. J. Phys. Chem. A, submitted.

20 Finlayson-Pitts, B. J.; Pitts, Jr., J. N. Chemistry of the Upper and Lower Atmosphere:

Theory, Experiments, and Applications. San Diego, CA : Academic Press, 2000.

21 Gaussian 98, Revision A.11.03, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria,

G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A., Jr.;

Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K.

249

N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci,

B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.;

Cui, Q.; Morokuma, K.; Salvador, P.; Dannenberg, J. J.; Malick, D. K.; Rabuck, A. D.;

Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Baboul, A. G.; Stefanov,

B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.;

Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.;

Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Andres, J. L.; Gonzalez, C.; Head-

Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian, Inc., Pittsburgh PA, 2001.

22 Gaussian 03, Revision C.02, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G.

E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.;

Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi,

M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.;

Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.;

Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;

Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;

Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.;

Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain,

M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.;

Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.;

Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham,

M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.;

250

Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A. Gaussian, Inc., Wallingford CT,

2004.

23 (a) Becke, A. D. Phys. Rev. A. 1988, 38, 3098-3100. (b) Lee, C.; Yang., W.; Parr, R.

G. Phys. Rev. B. 1988, 37, 785-789. (c) Becke, A. D. J. Chem. Phys. 1993, 98, 1372.

24 Bauschlichter, C. W., Jr.; Langhoff, S. R. Mol. Phys. 1999, 96, 471-476.

25 (a) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A., Jr. J. Chem. Phys. 1996,

104, 2598-2619; (b) Petersson, G. A.; Frisch, M. J. J. Phys. Chem. A 2000, 104, 2183.

26 Scott, A. P.; Radom, L. J. Phys. Chem. 1996, 100, 16502-16513.

27 (a) Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1989, 90, 2154. (b) Gonzalez, C.;

Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523.

28 (a) Reed, A. E.; Weinstock, R. B.; Weinhold, F. J. Chem. Phys. 1985, 83, 735-746. (b)

Reed, A. E; Curtiss, L. A.; Weinhold, F. Chem. Rev. 1988, 88, 899-926.

29 Hehre, W. J.; Radom, L.; Schleyer, P. v. R.; Pople, J. A. Ab Initio Molecular Orbital

Theory; John Wiley and Sons: New York, 1986.

30 Carter, A. Classical and statistical thermodynamics. Upper Saddle River: Prentice

Hall, 2001.

31 Using software designed by Dr. Timothy Barckholtz (Exxon-Mobil), obtained via

private communication.

32 Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6324-6331.

33 Chapter 3 of this work.

34 Liu, J.; Hadad, C. M.; Platz, M. S. Org. Lett. 2005, 7, 549-552.

251

CHAPTER 5

COMBUSTION PATHWAYS OF THE ALKYLATED HETEROAROMATICS: REACTIONS OF H, O (3P), AND HO• WITH THE ALKYLATED AZABENZENES

5.1 Introduction.

Coal is a valuable energetic resource; however, experimental and theoretical studies of this fossil fuel are limited due to its size and complexity. One common method for understanding the combustion reactions of coal involves modeling the reactions of the monocyclic aromatic heterocycles. Our previous work in this area has examined both non-alkylated and alkylated heteroaromatic rings; the latter has been proposed to best mimic the overall chemistry of the constituents of coal. We have studied the bond dissociation enthalpies1 and oxidation thermodynamics2 of the methyl- and ethyl-

substituted heteroaromatic rings; we have also explored the pathways that are available to

the resultant peroxy radicals of the alkylated heteroaromatics. So far, this work has been confined to calculating specific energies and enthalpies for reaction steps of interest; as such, comparisons with experiment have been limited.

Reactions of the alkylated heteroaromatic rings with reactive species such as hydroxyl radical, oxygen atom, and hydrogen atom are common to atmospheric oxidation and high-temperature combustion. Each of these reactive species is capable of either

252

abstracting a hydrogen atom from the parent molecule or adding to the aromatic system.

With substituted heteroaromatic rings, there are several non-equivalent sites where either

abstraction or addition can occur. For instance, for 2-methylpyridine, hydrogen-atom

abstraction can occur from the sp3 C−H of the methyl group or from the four unique sp2

C−H bonds of the ring itself. Moreover, a reactive species can add to any of the six positions on pyridine’s aromatic ring. Thus, eleven distinct pathways must be explored to obtain an overall picture of 2-methylpyridine’s reactivity; these pathways will be discussed in greater detail subsequently. It can be seen that heteroatom effects and alkyl substitution both play a role in increasing the complexity of the reactions of these heteroatomic rings; in the case of an alkylated aromatic hydrocarbon such as toluene, or of a non-alkylated heteroaromatic such as pyridine, several of these reaction pathways would be equivalent due to symmetry. No such simplification is possible for the vast majority of the alkylated heteroaromatics.

253

N N N

+ +

N N ABSTRACTION + + (+ H2O)

HO N

HO OH

ADDITION N N N + + OH

OH H H N N N + + + HO

H

OH H OH

Figure 5.1. Possible pathways for the reaction of hydroxyl radical (HO•) with 2- methylpyridine. Abstraction pathways are shown at the top, while radical-addition processes are shown at the bottom. Each abstraction pathway would generate water; only one equivalent is shown, for the sake of illustration.

254

The reactions of the alkylated heteroaromatics with reactive species have been

subjected to experimental analyses, notably in the case of the picolines (methyl-

substituted pyridines). Frerichs et al. monitored the reaction of the picolines with atomic

oxygen3 over the temperature range of 360−870 K, obtaining the following rate equations

(units of cm3/(molecule-s): for 2-picoline, k(T) = (5.43 x 10−11)exp[(−20.4 kJ/mol)/RT]; for 3-picoline, k(T) = (5.51 x 10−11)exp[(−20.7 kJ/mol)/RT; and for 4-picoline, k(T) =

(5.33 x 10−11)exp[(−20.4 kJ/mol)/RT]. Both the pre-exponential Arrhenius factors and

the overall activation energies were similar regardless of the location of the substituted

methyl group. Doughty and Mackie4 explored the reaction of 2-picoline with hydrogen

atom over the temperature range of 1300−1550 K and developed the rate expression: k(T)

= (1.66 x 10−11)exp[(−13.4 kJ/mol)/RT]. More recently, Yeung and Elrod used mass

spectrometric techniques and density functional theory (DFT)5 calculations to generate

rate expressions at 298 K and 100 Torr for the reactions of the picolines with hydroxyl

radical: for 2-picoline, k = 2.79 x 10−12; for 3-picoline, k = 2.3 x 10−12; and for 4-picoline,

k = 2.69 x 10−12 cm3/(molecule-s). This study also examined the rate coefficients for the

disubstituted methylpyridines (the lutidines) and the ethylpyridines. Overall, they noted an overall reactivity of: kpyridine < kpicolines < kethylpyridines < klutidines, suggesting that

increasing alkyl chain length and substitution decreased the reactivity of the azabenzenes.

As might be expected, studies on the reactivity of benzene and toluene with H,

O, and HO• are more widespread in the literature. These steps are key aspects of the

combustion mechanisms of benzene6 and toluene,7 which have been examined in several

255

studies. Specifically, the potential energy surface for the reaction of benzene and O (3P) has been a frequent target of both experimental8 and theoretical9 chemists. Most

recently,10 Nguyen et al. generated CBS-QB3 calculations of the potential energy surface

and used RRKM theory to determine that both the singlet and triplet forms of oxygen

atom have implications for the overall reaction of benzene and oxygen atom; they still

noted the dominance of the addition pathways but discounted the prevalence of the

cyclopentadiene + HCO• step that had been generally hypothesized to play a major role at higher temperatures. Phenoxy radical,11 as the benzene + O (3P) adduct, has been the

subject of considerable study. Addition pathways dominate for hydrogen atom and

hydroxyl radical, as well, at 298 K.12 Hydrogen atom prefers the addition pathway,

yielding cyclohexadienyl radical, at temperatures ranging up to 1000 K.13 Several

relevant studies of the reaction of HO• with benzene, toluene, and other unsaturated

systems have been compiled in review form by Atkinson et al.14,15 In general, these reactions differ as temperature increases: from 298-325 K, the reaction of benzene and

HO• generates hydroxycyclohexadienyl radical, which can be collisionally stabilized but

tends to decompose back to reactants;16 from 325-600 K, the hydroxycyclohexadienyl

product is dominant;17 and at higher temperatures (T > 600 K), hydrogen abstraction

(yielding phenyl radical and water) competes with a phenol generation pathway.18 The low temperature (298 K) reactions of HO• are of interest to atmospheric chemistry.

Previous work in our group by Barckholtz et al.19 used transition state theory

(TST) and DFT free energies to examine the reaction of several small aromatic

hydrocarbons and heterocycles: benzene, pyridine, furan, thiophene, and pyrrole. Via the

256

calculated free energies of activation and the corresponding rate coefficients derived from transition state theory, it was seen that radical addition was favored over radical abstraction at 298 K; as the temperature increased, H-atom abstraction became dominant due to the entropic advantage of such a process. The crossover temperature for the change in favor for H-atom abstraction vs. radical addition was seen to rely on the identity of both the aromatic species of interest and the reactive species [in terms of reactivity, O(3P) < H• < HO•)]. Both barrier height and abstraction-site preference varied linearly with the bond dissociation enthalpies (BDEs) for all species of interest. Radical addition occurred with lower barrier heights and greater exoergicities for the five- membered rings than for the six-membered rings, while hydrogen-atom abstraction occurred with lower barrier heights and greater exoergicities for the six-membered rings than for the five-membered rings.

Overall, it can be seen that several factors affect the reactivity of a given aromatic species, such as temperature, ring size, substitution patterns, and identity of the reactive species. This work examines some common trends that affect the reactivity of the alkylated azabenzenes (Figure 5.2a), including substitution patterns, reactive species identity, and the number of nitrogen atoms.

257

N N N N (a) 2 N 2 2

3 3 5 N N 4 4 4 Pyridine Pyridazine Pyrimidine Pyrazine

(b)

N

N N

Toluene 2-Picoline 3-Picoline 4-Picoline

Figure 5.2. (a) The azabenzenes. (b) The picolines (methylpyridines) and toluene, model compounds for understanding the chemistry of the azabenzenes with H•, O (3P), and HO•.

Barrier heights for each unique addition and abstraction reaction have been obtained, which will ultimately be used to explore the kinetics of these reactions and generate rate coefficients. It may be instructive to review how these quantities can ultimately be applied to exploring the kinetics of given processes. Reaction rate constants for a given reaction are calculated via the equation:

TS TQ )( BTk ≠ ( TST Γ= TTk )() exp Δ− 0 BT)/kG( (5.1) ring X TQTQ )()( h

≠ T is the temperature, h is Planck’s constant, kB B is the Boltzmann constant, ΔG 0 is the free energy of activation (barrier height) relative to reactants at infinite separation, and Q

258

refers to the partition function for each species involved in the reaction step (TS =

transition state, ring = aromatic ring, and X = reactive species). Finally, Γ(T) represents

quantum mechanical tunneling; in our group’s previous work, we have utilized the

Wigner approximation20 to tunneling:

2 1 ⎛ hvi ⎞ T 1)( +=Γ ⎜ ⎟ (5.2) 24 ⎝ BTk ⎠

where νi is the imaginary vibrational frequency corresponding to the transition state.

Once rate constants are calculated at various temperatures, the Arrhenius relationship can

be used to obtain the activation energy and other parameters of interest:

⎛ −E ⎞ ⎜ a ⎟ k(T) = Ae⎝ RT ⎠ (5.3)

⎛ −E ⎞ ⎜ o ⎟ k(T) = AT me⎝ RT ⎠ (5.4)

Equation 5.3 represents a temperature-independent energy of activation, while Equation

5.4 reflects a temperature-dependent activation barrier. The quantities A, Ea, E0, and m can all be obtained via regression analyses of k as a function of T (for instance, with equation 5.3, if a plot of ln k as a function of 1/T is generated, fitting the data yields a line with a slope equal to –Ea/R and a y-intercept of ln A).

The first step to obtaining these kinetic data is the identification of the transition

states for each potential reaction step. In this work, the enthalpies and energies of

activation and reaction at 298 K are reported for the reactions of the methyl-substituted

azabenzenes with H, O (3P), and HO•. The resulting energies of activation and reaction

are discussed as a function of various quantities, and overall trends are summarized.

259

5.2. Computational Methods

All geometry optimizations, vibrational frequency calculations, and single-point

energy calculations were completed with Gaussian 9821 and 0322 at the Ohio

Supercomputer Center. The B3LYP/6-31G* hybrid density functional theory (DFT)23 level was used for all geometry optimizations and vibrational frequency calculations.

Single-point energies were calculated at the B3LYP/6-311+G** level, using six Cartesian d functions with the scf=tight option; it has been shown that there is a small basis set effect with more polar systems,24 so this larger basis set was selected to compensate for

that effect. Vibrational frequencies were calculated for each stationary point to

characterize these structures as minima or transition states. The unscaled vibrational

frequencies were used to calculate the thermodynamic corrections to the enthalpy and free energy. Once obtained, zero-point vibrational energy corrections were scaled by a factor of 0.9806.25 The overall enthalpy at each temperature was determined from the

single-point energy, the thermal correction to the enthalpy, and the scaled zero-point energy, while the overall free energy at each temperature also included the entropic correction to the free energy.

Transition states were connected to reactants and products by either using

intrinsic reaction coordinate (IRC)26 searches or displacing the transition state geometries

by ±10% along the reaction coordinate characterized by the imaginary vibrational

frequency, then carefully optimizing (opt=calcfc) these resulting geometries to their

corresponding minima.

260

For the H and HO• systems, each species was treated as a doublet. For the doublet

radical species and many transition states, spin contamination () was negligible (no

greater than 0.80). The DFT methods employed in this work have demonstrated an

ability to minimize spin contamination for the radicals of interest, keeping the computed

values near the expected value of 0.75. For the O (3P) systems, the transition states

were treated as triplets.

In general, while B3LYP has been shown to perform well in a variety of

applications, generating barrier heights has been an area of concern for this method; some studies have demonstrated that it underestimates barrier heights, while others have shown that it is in good quantitative agreement with experimental activation energies and high-

level calculations. For the purposes of this work, we aim to obtain information on the

reactivity preferences of the alkylated heterocycles (i.e. the relative barrier heights for H-

atom abstraction and radical addition to the ring for the various parent molecules), rather

than directly obtain quantitatively correct rate coefficients; this is especially important as

we approach even larger systems of relevance to coal combustion. Further work might

involve geometries and energies obtained using Complete Basis Set (CBS-QB3)

methods,27 to validate our DFT approach for these particular systems.

261

5.3. Results and Discussion

5.3.1. Reactions of the Picolines.

For the methyl-substituted pyridines (Figure 5.2b), the free energies of activation and for the overall reaction are summarized (Tables 5.1−5.6) for abstraction and addition reactions with hydrogen atom (H•), oxygen atom [O (3P)], and hydroxyl (HO•) radical, along with the corresponding quantities for the hydrocarbon analogue, toluene. This initial work was undertaken to explore the effect that a nitrogen atom had on the reactivity of the aromatic ring system.

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline -- 16.38 16.25 17.25 13.39 9.23 3-picoline 13.47 -- 16.39 16.94 13.86 8.98 4-picoline 13.22 17.20 -- 17.20a 13.22 a 9.52 toluene 16.32 16.77 16.65 16.77 a 16.32 a 8.77 ADDITION N (1) 2 3 4 5 6 2-picoline 8.04 12.56 9.40 10.41 9.76 10.06 3-picoline 8.06 9.37 12.05 9.95 9.97 9.80 4-picoline 8.29 10.39 9.69 13.01 9.69 a 10.39 a toluene 12.00 9.33 9.77 9.61 9.77 a 9.33 a aPathway obtained via symmetry.

Table 5.1. Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for hydrogen-atom (H•) reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule. Free energies of activation are also included for toluene.

262

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline -- 3.98 3.47 3.58 -1.38 -16.78 3-picoline -1.77 -- 3.48 4.74 -0.67 -17.92 4-picoline -1.09 4.64 -- 4.64a -1.09 a -16.80 toluene 6.73 7.04 7.61 7.04 a 6.73 a -18.07 ADDITION N (1) 2 3 4 5 6 2-picoline -23.74 -12.99 -18.31 -15.49 -17.30 -16.57 3-picoline -24.36 -18.01 -14.24 -17.22 -16.99 -17.40 4-picoline -23.52 -15.97 -17.63 -12.08 -17.63 a -15.97 a toluene -13.55 -17.27 -16.12 -17.26 -16.12 -17.27 aPathway obtained via symmetry.

Table 5.2. Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for hydrogen-atom (H•) reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule. Free energies of reaction are also included for toluene.

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline - 11.61 10.66 11.88 7.13 3.25 3-picoline 7.74 - 11.10 11.34 7.57 4.53 4-picoline 7.44 12.48 - 12.48 a 7.44 a 6.24 toluene 11.37 11.42 10.59 11.42 a 11.37 a 3.72 ADDITION N (1) 2 3 4 5 6 2-picoline 13.34 11.49 6.73 9.49 6.73 8.76 3-picoline 13.03 8.14 8.01 8.55 7.47 8.22 4-picoline 13.25 9.32 7.05 10.47 7.05 a 9.32 a toluene 8.00 6.35 8.25 6.35 a 8.25 a 3.73 aPathway obtained via symmetry

Table 5.3. Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for oxygen atom [O (3P)] reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule. Free energies of activation are also included for toluene. All geometries involving oxygen addition were treated as triplet biradicals.

263

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline -- 5.26 4.76 4.87 -0.09 -15.49 3-picoline -0.49 -- 4.76 6.02 0.62 -16.63 4-picoline 0.20 5.92 -- 5.92a 0.20 a -15.52 toluene 9.87 10.19 10.75 10.19 a 9.87 a -16.79 ADDITION N (1) 2 3 4 5 6 2-picoline -1.15 -7.13 -10.41 -4.78 -9.32 -9.35 3-picoline -1.44 -10.58 -7.97 -6.80 -8.65 -10.35 4-picoline -1.15 -8.47 -9.48 -3.68 -9.48 a -8.47 a toluene -6.91 -8.68 -7.47 -8.98 -7.47 a -8.68 a aPathway obtained via symmetry

Table 5.4. Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for oxygen atom [O (3P)] reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule. Free energies of reaction are also included for toluene. All geometries involving oxygen addition were treated as triplet biradicals.

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline - 8.47 7.73 8.21 4.75 5.41 3-picoline 5.55 - 8.40 7.95 5.07 4.22 4-picoline 4.69 9.16 - 9.16 a 4.69 a 4.15 toluene 7.85 8.88 7.97 8.88 7.85 3.54 ADDITION N (1) 2 3 4 5 6 2-picoline 21.41 10.94 7.16 9.96 7.31 8.95 3-picoline 20.38 7.41 8.26 8.83 7.93 8.46 4-picoline 21.01 9.42 7.40 10.66 7.40 a 9.42 a toluene 7.77 6.70 7.45 6.71 7.45 a 6.70 a .aPathway obtained via symmetry

Table 5.5. Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for hydroxyl radical (HO•) reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule. Free energies of activation also included for toluene.

264

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline - -5.97 -6.47 -6.36 -11.32 -26.72 3-picoline -11.72 - -6.47 -5.20 -10.61 -27.86 4-picoline -11.03 -5.31 - -5.31 a -11.03 a -26.75 toluene -4.95 -4.64 -4.08 -4.64 -4.95 -28.01 ADDITION N (1) 2 3 4 5 6 2-picoline 21.53 -1.05 -3.95 -0.08 -3.16 -7.15 3-picoline 19.89 -8.68 -2.24 -1.62 -2.62 -8.17 4-picoline 20.99 -6.32 -3.37 1.23 -3.37 a -6.32 a toluene -1.54 -3.32 -2.21 -3.41 -2.21 a -3.32 a aPathway obtained via symmetry

Table 5.6. Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for hydroxyl radical (HO•) reactions with the picolines (methyl-substituted pyridines), at each relevant site within the parent molecule. Free energies of reaction are also included for toluene.

The presence of a nitrogen atom does affect the chemistry of the aromatic system,

usually by a quantitatively small amount (~2 kcal/mol). Notably, the abstraction

pathways are seen to proceed with lower kinetic barriers for the picolines than for toluene; this is attributed to the ability of the nitrogen atom to provide greater resonance stability within the resulting picolinyl radicals,28 correspondingly stabilizing the

transition states leading to these radicals. The lone exception occurs for the abstraction

of a methyl hydrogen, which is generally more favorable for toluene than the picolines, regardless of position of abstracting species. This correlates with our work on the bond dissociation enthalpies (BDEs) of the alkylated heteroaromatics: it was seen that the picolinyl radicals cannot delocalize their electron density to as great as an extent as benzyl radical, so these radicals are less stable and their methyl C−H BDEs are

265

correspondingly higher. For the addition pathways, toluene demonstrates substantially

lower barriers to addition to the ring than did the picolines; this is in keeping with similar

previous studies.

In terms of thermodynamics, both abstraction from and addition to the alkylated

heteroaromatic rings are more favorable than the corresponding reactions of toluene. For all of the compounds explored, the addition pathways were calculated to be exoergic, slightly moreso for the picolines. The abstraction pathways of the picolines are slightly endoergic overall (ΔGo < 5 kcal/mol); with toluene, abstraction of an aromatic C−H bond

generally costs approximately 10 kcal/mol. In all cases, abstraction of the methyl

hydrogen is kinetically competitive and thermodynamically dominant. This was unsurprising, since this route maintains the aromaticity of the ring for all compounds and generates stable abstraction products. With respect to the latter consideration, reactivity

3 • increases for all the alkylated heteroaromatics, via the trend H < O ( P) < HO , as H2,

HO•, and water are generated as products, respectively.

Besides the strong reactive preference for abstraction of hydrogen atom from the

methyl group, a few other trends can be discerned for the addition reactions. Addition at

N invokes a high kinetic barrier for HO• and O, but is a relatively reactive site for H-atom addition; the respective energies of reaction match this trend. Kinetic barriers and energies of reaction for addition to the other sites are competitive; only a few slight trends can be discerned. It appears that reaction with HO• favors addition ortho to the

nitrogen atom, likely due to the resonance effects provided by the heteroatom; similarly,

addition ortho to the position of methyl substitution is a preference seen for O and H:

266

increasing alkyl substitution would be afforded to the incipient radical via these routes.

With respect to the addition reactions of toluene, reaction is consistently favored at the positions ortho and para to the methyl group, which would invoke a similar stabilizing effect on the resulting adduct.

The identity of the reactive species does play a role in both the overall reactivity

and the addition pathways favored; the trend of increasing kinetic reactivity for H ~ O

(3P) < HO• is consistently seen across the possible abstraction and addition pathways, as is the thermodynamic preference of O (3P) < H < HO•. Such preferences have been

demonstrated in similar work. Hydrogen-atom abstraction from the substituted methyl

group is consistently favored regardless of aromatic or reactive species; however, beyond

this fact, several variations in reaction pathways can be seen. Hydrogen-atom abstraction

is the preferred mechanism of HO• radical at 298 K, while H atom favors addition to the

ring, and addition and abstraction are competitive for O (3P). As shown, these

differences in addition preferences have minor effects at 298 K, but it is likely that these

ramifications will be minimized at higher temperatures (of interest for flame

combustion), due to the entropic advantage of the abstraction processes. We also note

here that our treatment of the oxygen atom was incomplete; for simplicity’s sake, the

resulting species were treated as biradicals, so only the triplet states were considered for

reactants, transition state, and products, in these reactions. To fully understand the scope of these reactions, it would be necessary to incorporate multiple spin possibilities in generating the potential energy surfaces for these reactions. However, it is hoped that the

qualitative comparisons obtained via these elementary calculations may still prove useful.

267

The effect of alkylation on the relative preferences for these pathways could be explored, since previous work in our group has calculated the barriers to reaction for H-

atom abstraction and addition to benzene and pyridine, the non-alkylated equivalents of

these species. In this case, both abstraction and addition were generally more favorable

for the hydrocarbon than the azabenzyl equivalent (Tables 5.7 and 5.8). It can be seen

that increasing alkyl substitution increases the barrier heights for these reactions,

suggesting that steric strain is not negligible in the transition states. Also, as mentioned

previously, the picolines consistently demonstrate lower barriers to hydrogen-atom

abstraction than does toluene, whereas pyridine was shown to be consistently less

reactive to hydrogen-atom abstraction than was benzene. Alkyl substitution does result

in minor qualitative and quantitative changes to the reactivity of these species.

268

Abstraction Addition

Site H O (3P) OH H O (3P) OH

Benzene 1 13.64 10.10 7.20 8.28 6.36 6.73

Pyridine 2 13.25 7.70 5.13 9.75 9.94 10.22

3 16.66 11.97 8.35 9.58 7.83 8.18

4 16.36 9.44 8.09 9.58 9.91 10.04

Table 5.7. Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for the H-atom abstraction pathways and addition pathways available to benzene and pyridine. Data obtained from Ref. 19.

Abstraction Addition

Site H O (3P) OH H O (3P) OH

Benzene 1 3.46 4.97 -5.92 -17.42 -3.06 -5.26

Pyridine 2 -1.52 -0.01 -10.90 -17.20 -8.93 -6.95

3 4.54 6.05 -4.84 -17.69 -8.70 -4.25

4 3.67 5.19 -5.70 -16.16 -5.18 -2.12

Table 5.8. Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for the H-atom abstraction pathways and addition pathways available to benzene and pyridine. Data obtained from Ref. 19.

269

5.3.2. Reactions of the Alkylated Diaazabenzenes

The effect of multiple heteroatoms on reactivity with O (3P) and HO• was also

explored, by comparison of the energetics of the picolines with those of the methyl-

substituted diazabenzenes. The energies of activation and reaction are compiled (Tables

5.9−5.12) for the reactions of O (3P) and HO• with 2-, 4-, and 5-methylpyrimidine; 3- and

4-methylpyridazine, and 2-methylpyrazine.

ABSTRACTION 2 3 4 5 6 methyl 2-picoline - 11.61 10.66 11.88 7.13 3.25 3-picoline 7.74 - 11.10 11.34 7.57 4.53 4-picoline 7.44 12.48 - 12.48 a 7.44 a 6.24 2-methylpyrimidine - - 7.36 14.32 7.36 a b 4-methylpyrimidine 5.95 - - 9.26 7.35 2.50 5-methylpyrimidine b - 8.54 - 8.54 a 6.53 3-methylpyridazine - - 9.03 9.08 6.15 4-methylpyridazine - 9.82 - 11.69 9.17 6.65 2-methylpyrazine - 7.90 - 7.99 7.42 5.35 ADDITION N 2 3 4 5 6 2-picoline 13.34 11.49 6.73 9.49 6.73 8.76 3-picoline 13.03 8.14 8.01 8.55 7.47 8.22 4-picoline 13.25 9.32 7.05 10.47 7.05 a 9.32 a 2-methylpyrimidine 14.20 13.42 14.20 12.02 7.31 12.02 a 4-methylpyrimidine 10.01 7.28 10.15 8.56 3.42 7.34 5-methylpyrimidine 14.37 11.25 14.37a 10.75 9.24 10.75 a 3-methylpyridazine 8.91 9.48 12.81 8.66 9.70 10.82 4-methylpyridazine 9.83 9.54 11.21 11.41 9.17 11.81 2-methylpyrazine 12.79 9.77 8.08 12.69 8.08 8.46 a Equivalent to previous path via symmetry. b Optimization consistently returned second-order saddle-point rather than TS.

Table 5.9. Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reactions of oxygen atom [O (3P)] and the methyl-substituted azabenzenes, at each relevant site within the parent molecule.

270

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline -- 5.26 4.76 4.87 -0.09 -15.49 3-picoline -0.49 -- 4.76 6.02 0.62 -16.63 4-picoline 0.20 5.92 -- 5.92a 0.20 a -15.52 2-methylpyrimidine -- -- -0.98 7.68 -0.98 a -13.86 4-methylpyrimidine -2.63 -- -- 2.15 -4.62 -18.62 5-methylpyrimidine 2.52 -- -0.48 -- -0.48 a -11.63 3-methylpyridazine -- -- 2.93 3.39 1.87 -15.00 4-methylpyridazine -- 1.22 -- 4.00 1.85 -15.74 2-methylpyrazine -- -0.68 -- 0.82 -0.03 -15.78 ADDITION N 2 3 4 5 6 2-picoline -1.15 -7.13 -10.41 -4.78 -9.32 -9.35 3-picoline -1.44 -10.58 -7.97 -6.80 -8.65 -10.35 4-picoline -1.15 -8.47 -9.48 -3.68 -9.48 a -8.47 a 2-methylpyrimidine -1.21 b -1.21 a -5.82 -8.95 -5.82 a 4-methylpyrimidine -4.73 -9.81 -4.64 -7.48 -13.78 -9.76 5-methylpyrimidine -0.58 -7.53 -0.58 a -7.59 -7.95 -7.59 a 3-methylpyridazine -11.92 -11.58 b -10.99 -8.48 -36.95 4-methylpyridazine -11.10 -10.54 -39.64 -7.56 -10.54 -36.81 2-methylpyrazine -1.29 -11.10 -15.02 0.94 -13.81 -13.12 aPathway obtained via symmetry. bGeometry could not be optimized.

Table 5.10. Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reactions of oxygen atom [O (3P)] and the methyl-substituted azabenzenes, at each relevant site within the parent molecule.

271

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline - 8.47 7.73 8.21 4.75 5.41 3-picoline 5.55 - 8.40 7.95 5.07 4.22 4-picoline 4.69 9.16 - 9.16 a 4.69 a 4.15 2-methylpyrimidine - - 4.40 9.12 4.40 a 6.72 4-methylpyrimidine 2.26 - - 5.29 0.44 2.14 5-methylpyrimidine 7.27 - 6.06 - 6.06 a 4.94 3-methylpyridazine - - 6.39 6.26 6.24 5.76 4-methylpyridazine - 7.01 - 7.65 6.25 5.58 2-methylpyrazine - 5.32 - 5.10 4.67 5.71 ADDITION N 2 3 4 5 6 2-picoline 21.41 10.94 7.16 9.96 7.31 8.95 3-picoline 20.38 7.41 8.26 8.83 7.93 8.46 4-picoline 21.01 9.42 7.40 10.66 7.40 a 9.42 a 2-methylpyrimidine 19.43 15.01 19.43 a 12.46 14.73 12.46 a 4-methylpyrimidine 17.87 8.25 18.26 11.09 3.72 7.98 5-methylpyrimidine 18.92 12.37 18.92 a 10.04 9.30 10.04 a 3-methylpyridazine 8.93 9.70 11.32 8.58 9.50 9.29 4-methylpyridazine 10.25 9.68 8.73 10.35 8.95 10.23 2-methylpyrazine 16.08 10.60 7.41 18.25 9.09 9.20 aPathway obtained via symmetry.

Table 5.11. Free energies of activation (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reactions of hydroxyl radical (HO•) and the methyl-substituted azabenzenes, at each relevant site within the parent molecule.

272

ABSTRACTION 2 3 4 5 6 Methyl 2-picoline - -5.97 -6.47 -6.36 -11.32 -26.72 3-picoline -11.72 - -6.47 -5.20 -10.61 -27.86 4-picoline -11.03 -5.31 - -5.31 a -11.03 a -26.75 2-methylpyrimidine - - -12.21 -3.55 -12.21 a -25.09 4-methylpyrimidine -13.87 - - -9.08 -15.85 -29.85 5-methylpyrimidine -8.71 - -11.71 - -11.71 a -22.86 3-methylpyridazine - - -8.30 -7.84 -9.36 -26.23 4-methylpyridazine - -10.01 - -7.23 -9.38 -26.97 2-methylpyrazine - -11.91 - -10.41 -11.26 -27.01 ADDITION N 2 3 4 5 6 2-picoline 21.53 -1.05 -3.95 -0.08 -3.16 -7.15 3-picoline 19.89 -8.68 -2.24 -1.62 -2.62 -8.17 4-picoline 20.99 -6.32 -3.37 1.23 -3.37 a -6.32 a 2-methylpyrimidine 19.95 -2.46 19.95 a -3.69 -4.12 -3.69 a 4-methylpyrimidine 17.62 -8.77 18.17 -1.82 -8.01 -8.12 5-methylpyrimidine 19.51 -5.92 19.51 a -6.23 -2.07 -6.23 a 3-methylpyridazine 5.35 6.22 -5.31 -5.18 -2.72 -7.51 4-methylpyridazine 7.69 6.32 -7.67 -1.51 -3.81 -5.69 2-methylpyrazine 16.46 -7.01 -10.41 16.31 -9.40 -8.74 aPathway obtained via symmetry.

Table 5.12. Free energies of reaction (B3LYP/6-311+G**//B3LYP/6-31G*, kcal/mol, 298K) for reactions of hydroxyl radical (HO•) and the methyl-substituted azabenzenes, at each relevant site within the parent molecule.

The presence of a second nitrogen atom appears to stabilize the radicals formed

via hydrogen-atom abstraction processes, as the abstraction products of the methyl-

substituted diazabenzenes consistently demonstrate lower barrier heights and greater

exoergicities than those seen with the picolines. No major changes are seen in the

addition products: addition to nitrogen still occurs with a substantial barrier, and addition

to the positions ortho to nitrogen generally invokes the lowest barriers, closely followed by addition at those positions ortho to the methyl group.

273

The addition reactions of the methylpyridazines with O (3P) at positions 3 and 6 demonstrate considerable exoergicity, and lead to a ring-opened reaction product from the relevant transition states, due to the ability of these species to form diazo- functionalized carbonyl compounds (Figure 5.3).

ν =482i

Figure 5.3. Representation of ring-opening reactions unique to methylpyridazines, when atomic oxygen attacks ortho to the nitrogen atoms. Positive displacement (10%) along the reaction coordinate characterized by the imaginary frequency associated with the transition states of the reactions of the methylpyridazines and O (3P) yields ring-opened products. Reaction shown for 3-methylpyridazine with O (3P) adding to position 3. Similar reactions occur for 4-methylpyridazine with O (3P) adding to position 3, 3- methylpyridazine with O (3P) adding to position 6, and 4-methylpyridazine with O(3P) adding to position 6.

Such reactions are possible whenever oxygen atom adds ortho to either of the two nitrogen atoms of the pyridazines. These geometries will be explored further; while extreme care was taken in generating the relevant minima, the energy difference between these adducts and others arising from O (3P) addition is such that the ring-opening

processes would have an enormous influence on subsequent kinetic calculations, so it is

important to ensure that these are indeed the relevant products from displacement of these

274

transition states. The reaction coordinate will be explored to check if isolable minima

exist for the ring-closed products.

Overall, the picolines provided a good benchmark for the overall reactivity of the

methylated azabenzenes with O (3P) and HO•. The key difference was a small increase in

stability of the abstraction products for the diazabenzenes, presumably due to the greater

resonance effects generated from a second heteroatom. When all of the relevant energies

of activation and reaction were considered, it was seen that radical addition to an

aromatic ring carbon was generally the more favorable process at 298 K for O (3P), while abstraction was favored overall at 298 K for HO•. Addition to nitrogen atoms was consistently unfavorable relative to other potential pathways of these compounds.

Methyl abstraction was clearly the dominant step for both HO• and O (3P), occurring with

barrier heights < 5 kcal/mol and exoergicities on the order of –25 to –30 kcal/mol for the alkylated azabenzenes.

5.3.3. Overall trends

The correlation between the energies of abstraction for the alkylated azabenzenes

and the bond dissociation enthalpies of these species were explored, for both HO• (Figure

5.4) and O (3P) (Figure 5.5). Additionally, the correlation between barrier heights for

abstraction reactions and barrier heights for addition reactions was modeled (Figure 5.6),

at the relevant positions of the methyl derivatives of the azabenzenes.

275

Figure 5.4. Correlation between the bond dissociation enthalpies (kcal/mol) for hydrogen-atom loss at the methyl group of the alkylated heteroaromatic rings and the corresponding free energy of activation (kcal/mol) for H-atom abstraction at the methyl group by HO•.

276

Figure 5.5. Correlation between the bond dissociation enthalpies (kcal/mol) for hydrogen-atom loss at the methyl group of the alkylated heteroaromatic rings and the corresponding free energy of activation (kcal/mol) for H-atom abstraction at the methyl group by O (3P).

277

Figure 5.6. Relationship between the free energy of activation (kcal/mol) for H-atom abstraction by HO• at a given position and the free energy of activation (kcal/mol) for addition of HO• to that same position.

In general, the BDE values calculated for the methyl derivatives of the

azabenzenes1 correlated well with the activation barriers for H-atom abstraction at the methyl group by HO•. However, no real correlation was seen between the BDE values

and the activation barriers for H-atom abstraction by O (3P). This was not surprising,

given the differences in reactivity between HO• and O (3P), as well as the previously-

discussed drawbacks of our treatment of the spin system of the reactions involving

oxygen atom. Additionally, no correlation exists between the activation barrier for

278

abstraction at a given position and the activation barrier for addition at that same position; therefore, different factors affect the barrier heights for these types of reactions.

5.4. Conclusions

The kinetics and thermodynamics of the reactions of the methyl azabenzenes with

H, O (3P), and HO• were explored at 298 K using the B3LYP method. These alkylated heteroaromatics are of interest due to their ability to predict the chemistry associated with coal. It was seen that the alkylated azabenzenes demonstrate lower barriers to H-atom abstraction than their hydrocarbon analogue (toluene); these species also react more readily than the non-alkylated azabenzenes with respect to both abstraction and addition.

Regardless of reactive species or starting methyl azabenzene, the most favorable reaction for these species was abstraction of a methyl hydrogen, which led to preservation of the aromatic ring system (and increased delocalization onto the benzyl position). Other trends were also noted: at 298 K, HO• reacted preferentially via H-atom abstraction, while H atom tended to add to the aromatic ring, and O (3P) could undergo both abstraction and addition reactions at comparable energetic expense. The methyl derivatives of both mono- and di-azasubstituted heteroaromatic rings were explored; the key difference between these systems involved a greater preference for H-atom abstraction reactions in the methyl diazabenzenes, attributed to greater resonance stabilization caused by the second heteroatom. In general, the subset of the monoaza- heteroaromatics (picolines) predicted the major trends associated with all of the alkylated azabenzenes.

279

With respect to predicting the chemistry of the larger heteroaromatic systems

found in coal, these calculations imply that HO• will be the main species to initiate reaction. Since HO• tends to preferentially abstract H-atom from the alkylated

heteroaromatics even at the low temperature examined in this work (298 K), we

hypothesize that this pathway will quickly become significant as the entropic benefits of

abstraction become more pronounced as temperatures rise to a level of interest in high-

temperature combustion. Moreover, this work suggests that the combination of

increasing N-substitution in the aromatic ring and increasing alkyl substitution in a given

sample may lead to increasing overall H-atom abstraction reactivity with HO•.

The complicated structure of the alkylated heteroaromatic rings suggests that

modeling their kinetics will pose several challenges: as the size of a given system

increases, it becomes exponentially more difficult to incorporate all the possible

pathways necessary to accurately replicate the system’s combustion. In this work,

several potential elementary reaction steps of the methyl azabenzenes with common

flame intermediates have been systematically delineated. The activation barriers

calculated in this work can ultimately be used to generate rate coefficients; a series of rate

coefficients generated over a temperature range can be used to obtain the Arrhenius parameters for a given system; and these parameters can be used in the master equation methods that serve to model oxidation processes of complicated systems.

280

References for Chapter 5

1 Chapter 2 of this thesis.

2 Chapters 3 and 4 of this thesis.

3 Frerichs, H.; Schliephater, V.; Tappe, M.; Wagner, H. G. Z. Phys. Chem. 1990, 166,

157-165.

4 Doughty, A.; Mackie, J. C. J. Phys. Chem. 1992, 96, 10339-10348.

5 (a) Parr, R. G.; Yang, W. Density Functional Theory in Atoms and Molecules; Oxford

University Press: New York, 1989. (b) Labanowski, J. W.; Andzelm, J. Density

Functional Methods in Chemistry; Springer: New York, 1991.

6 (a) Asaba, T.; Fujii, N. Symp. (Int.) Combust., [Proc.] 14th, 1971, 155. (b) Venkat, C.;

Brezinsky, K.; Glassman, I. Symp. (Int.) Combust., 19th 1982, 143-152. (c) Bittker, D. A.

Prog. Energy Combust. Sci. 1991, 79, 49-72. (d) Zhang, H.-Y.; McKinnon, J. T.

Combust. Sci. Technol. 1995, 107, 261-300. (e) Liu, R.; Morokuma, K.; Mebel, A.M.;

Lin, M.C. J. Phys. Chem. 1996, 100, 9314-9322. (f) Olivella, S.; Sole, A.; Garcia-Raso,

A. J. Phys. Chem. 1995, 99, 10549-10556. (g) Hodgson, D.; Zhang, H.-Y.; Nimlos, M.

R.; McKinnon, J. T. J. Phys. Chem. A 2001, 105, 4316-4327. (h) Sethuraman, S.;

Senkan, S. M.; Gutman, D. Combust. Sci. Technol. 1992, 82, 13-30. (i) Carpenter, B. K.

J. Am. Chem. Soc. 1993, 115, 9806-9807.

7 (a) Emdee, J. L.; Brezinsky, K.; Glassman, I. J. Phys. Chem. 1992, 96, 2151–2161. (b)

Dagaut, P.; Pengloan, G.; Ristori, A. Phys. Chem. Chem. Phys. 2002, 4, 1846-1854. (c)

Pitz, W. J.; Seiser, R.; Bozzelli, J.W.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin-

Leclerc, F.; Seshadri, K.; Westbrook, C.K. 2nd Joint Meeting of the U.S. Section of the

281

Combustion Institute: Oakland, CA. March 25-28, 2001. (d) Pitz, W.J.; Seiser, R.;

Bozzelli, J.W.; Seshadri, K.; Chen, C. J.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin-

Leclerc, F.; Westbrook, C. K. Thirtieth Symposium (International) on Combustion. The

Combustion Institute; Chicago, IL: July 25-30, 2004. (e) El Bakali, A.; Dupont, L.;

Lefort, B.; Lamoureux, N.; Pauwels, J. F.; Montero, M. J. Phys. Chem. A 2007, 111,

3907-3921.

8 (a) Sibener, S. J.; Buss, R. J.; Casavecchia, P.; Hirooka, T.; Lee, Y. T. J. Chem. Phys.

1980, 72, 4341. (b) Bajaj, P. N.; Fontijn, A. Combust. Flame 1996, 105, 239.

9 (c) Hodgson, D.; Zhang, H. Y.; Nimlos, M. R.; McKinnon, J. T. J. Phys. Chem. A 2001,

105, 4316. (d) Ko, T.; Adusei, G. Y.; Fontijn, A. J. Phys. Chem. 1991, 95, 8745. (e)

Barry, N. J.; Fletcher, I. W.; Whitehead, J. C. J. Phys. Chem. 1986, 90, 4911. (f)

Avramenko, L. I.; Kolesnikova, R. V.; Savinova, G. I. Bull. Acad. Sci. U.S.S.R. Div.

Chem. Sci. (Engl. Transl.) 1965, 24-29. (g) Colussi, A. J.; Singleton, D. L.; Irwin, R. S.;

Cvetanovic, R. J. J. Phys. Chem. 1975, 79, 1900. (h) Atkinson, R.; Pitts, J. N. Chem.

Phys. Lett. 1979, 63, 485. (i) Cvetanovic, R. J. J. Phys. Chem. Ref. Data 1987, 16, 261.

(j) Leidreiter, H. I.; Wagner, H. G. Z. Phys. Chem. 1989, 165, 1. (k) Tappe, M.;

Schliephake, V.; Wagner, H. G. Z. Phys. Chem. 1989, 162, 129.

10 Nguyen, T. L.; Peeters, J.; Vereecken, L. J. Phys. Chem. A. 2007, 111, 3836-3849.

11 (a) Hausmann, M.; Hebgen, P.; Homann, K. H. Symp. (Int.) Combust., [Proc.] 24th,

1992, 793-801. (b) Wang, Z.; Wang, J.; Richter, H.; Howard, J. B.; Carlson, J.;

Levendis, Y. A. Energy Fuels, 2003, 17, 999-1013. (c) Sebbar, N.; Bockhorn, H.;

Bozzelli, J. W. J. Phys. Chem. A 2005, 109, 2233-2253; (d) Sebbar, N.; Bockhorn, H.;

282

Bozzelli, J. W. Phys. Chem. Chem. Phys. 2002, 15, 3691-3703; (e) Vovelle, C.; Delfau, J.

L. NATO Science Series, Series C: Mathematical and Physical Sciences 2000, 183-203.

12 (a) Atkinson, R. J. Phys. Chem. Ref. Data, Monograph 1, 1989, 1; (b) Atkinson, R. J.

Phys. Chem. Ref. Data, Monograph 2, 1994, 1.

13 (a) Mebel, A. M.; Lin, M. C.; Yu, T.; Morokuma, K. J. Phys. Chem. A 1997, 101,

3189-3196. (b) Kiefer, J. H.; Mizerka, L. J.; Patel, M. R.; Wei, H. C. J. Phys. Chem.

1985, 89, 2013-2019. (c) Rao, V. S.; Skinner, G. B. J. Phys. Chem. 1984, 88, 5990-5995.

(d) Asaba, T.; Fujii, N. Proc. 13th Int. Combust. Symp. 1971, 155-164. (e) Park, J.;

Dyakov, I. V.; Lin, M. C. J. Phys. Chem. A 1997, 101, 8839-8843. (f) Sauer, M. C., Jr.;

Ward, B. J. Phys. Chem. 1967, 71, 3971-3983. (g) Sauer, M. C., Jr.; Mani, I. J. Phys.

Chem. 1970, 74, 59-63. (h) Yang, K. J. Am. Chem. Soc. 1962, 84, 3795. (i) Knutti, R.;

Buhler, R. E. Chem. Phys. 1975, 7, 229-243. (j) Kim, P.; Lee, J. H.; Bonanno, R. J.;

Timmons, R. B. J. Chem. Phys. 1973, 59, 4593-4601. (k) Nicovich, J. M.; Ravishankara,

A. R. J. Phys. Chem. 1984, 88, 2534-2541.

14 Atkinson, R. Atmos. Environ. A 1990, 24A, 1-41.

15 Atkinson, R.; Arey, J. Chem. Rev. 2003, 103, 4605-4638.

16 (a) Davis, D. D.; Bollinger, W.; Fischer, S. J. Phys. Chem. 1975, 79, 293-294. (b) 43.

Wahner, A.; Zetsch, C. J. Phys. Chem. 1983, 87, 4945-4951.

17 (a) Perry, R. A.; Atkinson, R.; Pitts, J. N., Jr. J. Phys. Chem. 1977, 81, 296-304. (b)

Tully, F. P.; Ravishankara, A. R.; Thompson, R. L.; Nicovich, J. M.; Shah, R. C.;

Kreutter, N. M.; Wine, P. H. J. Phys. Chem. 1981, 85, 2262-2269. (c) Madronich, S.;

Felder, W. J. Phys. Chem. 1985, 89, 3556-3561.

283

18 Lin, C. Y.; Lin, M. C. J. Phys. Chem. 1986, 90, 425-431.

19 Barckholtz, C.; Barckholtz, T. A.; Hadad, C. M. J. Phys. Chem. A 2001, 105, 140-152.

20 Wigner, E. P. Z. Phys. Chem. Abst. B 1932, 19, 203.)

21 Gaussian 98, Revision A.11.03, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria,

G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A., Jr.;

Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K.

N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci,

B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.;

Cui, Q.; Morokuma, K.; Salvador, P.; Dannenberg, J. J.; Malick, D. K.; Rabuck, A. D.;

Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Baboul, A. G.; Stefanov,

B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.;

Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.;

Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Andres, J. L.; Gonzalez, C.; Head-

Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian, Inc., Pittsburgh PA, 2001.

22 Gaussian 03, Revision C.02, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G.

E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.;

Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi,

M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.;

Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.;

Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;

Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;

Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.;

284

Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain,

M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.;

Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.;

Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham,

M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.;

Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A. Gaussian, Inc., Wallingford CT,

2004.

23 (a) Becke, A. D. Phys. Rev. A. 1988, 38, 3098-3100. (b) Lee, C.; Yang., W.; Parr, R.

G. Phys. Rev. B. 1988, 37, 785-789. (c) Becke, A. D. J. Chem. Phys. 1993, 98, 1372-

1377.

24 Bauschlichter, C. W., Jr.; Langhoff, S. R. Mol. Phys. 1999, 96, 471-476.

25 Scott, A. P.; Radom, L. J. Phys. Chem. 1996, 100, 16502-16513.

26 (a) Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1989, 90, 2154. (b) Gonzalez, C.;

Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523.

27 (a) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A., Jr. J. Chem. Phys. 1996,

104, 2598-2619; (b) Petersson, G. A.; Frisch, M. J. J. Phys. Chem. A 2000, 104, 2183-

2190.

28 Barckholtz, C.; Barckholtz, T. A.; Hadad, C. M. J. Am. Chem. Soc. 1999, 121, 491-

500.

285

CHAPTER 6

THE MULTIPLE CONFORMATIONS OF N-PROPYLPEROXY RADICAL AND THEIR IMPLICATIONS FOR ITS UNIMOLECULAR DECOMPOSITION

Reproduced in part from J. Phys. Chem. A 2005, 109, 3637-3646. Copyright 2005,

American Chemical Society.

6.1 Introduction

Peroxy radicals constitute an important class of reactive oxygen species. Formed via the addition of molecular oxygen to a wide variety of alkyl and aromatic radicals, these reactive intermediates greatly affect atmospheric chemistry and also have been shown to influence high-temperature combustion. These species commonly react via rearrangements, H-atom abstractions, loss of oxygen, or β-scission.1

Ethylperoxy and propylperoxy radicals are generally used to model the chemistry of larger alkylperoxy species. Ethylperoxy radical is the predominant reactive radical intermediate in the low-temperature combustion of ethyl radical. It also is the smallest species wherein an internal H-atom transfer can feasibly occur. In a peroxy radical, an internal H-atom transfer yields a hydroperoxyalkyl radical, which can then be oxidized at the incipient radical center to give a hydroperoxyalkylperoxy radical. These processes are key in the chain-branching steps that lead to autoignition. Ignatyev et al. first

286

demonstrated that the unimolecular decomposition step for ethylperoxy radical likely

involves a concerted 1,4-H-atom transfer and hydroperoxyl elimination.2 This step and

other possibilities have been explored experimentally and modeled theoretically, as

compiled in the thorough review by Rienstra-Kiracofe et al.3

n-Propylperoxy radical can feasibly undergo several unimolecular decomposition steps, due to its chain length (Figure 6.1).

H O O

Figure 6.1. n-Propylperoxy radical. Note particularly that 1,5-H atom transfer can readily occur via a six-membered ring transition state.

Experimental studies4 have shown that propene and hydroperoxyl radical are the most common products, suggesting that 1,4-H atom transfer again plays a major role, whether it results in an isolable radical or is concerted with HO2 elimination. Previous

computational studies have modeled other H-atom transfers available to this species;

while 1,4-H atom transfer leads to the empirically-observed products, 1,5-H atom transfer

occurs via a stable six-membered transition state, so that it could also lead to feasible

unimolecular decomposition products. Both of these H-atom transfers saw a lower

barrier than that which was necessary to reform propyl radical and O2; n-propylperoxy

radical was a significant intermediate.

287

One concern that is not immediately evident with the chemistry of the alkylperoxy

species is their ability to exist in several conformations; this is an area that has been

largely overlooked by the qualitative studies previously discussed. Because these

conformations would be comparable in energy, it is presumed that they would be able to

interconvert readily. Several spectroscopic methods are capable of observing these non-

degenerate rotamers; the presence of multiple conformers has implications for

experimental findings in these areas. Moreover, these internal rotations also affect the

kinetics of the pathways available to applicable species. When a species can exist in

several conformeric forms, each of these forms must be considered in defining its overall

properties, such as dipole moment, vibration frequencies, moment of inertia, and rate of

reaction. In particular, rates of reaction involve classification of the internal rotations of a

species. Equation (6.1) shows the equation for calculating the rate coefficient of a given

process from the activation barrier (Eo), a tunneling correction Γ(T), and the total

‡ partition functions (QREACTANT, Q TS) for both the reactant and the transition state:

‡ −Eo Tk TS TQ )( k(T) = Γ(T) B e RT (6.1) Nh REACTANT TQ )(

Since the partition function for a given species depends in part on the degeneracy of that

species, it is important to account for the conformeric possibilities for both the reactant

and transition state. This is particularly important when a given reaction step (e.g. cyclization) involves a change in the number of conformers between reactant and transition state. Vereecken and Peeters have devoted significant time to studies of this nature, determining multirotamer rate coefficients for processes of the β-hydroxyalkoxy

288

radicals,5 as well as for the 1,5-H-atom shift in butoxy radical;6 they demonstrated that

accounting for all conformers in a kinetic study provided more accurate results than did

working with the lowest-energy conformer alone. The 1,5-H-atom shift in butoxy radical

is analogous to that seen in n-propylperoxy radical, so it seems that an understanding of

the conformeric distribution in the latter species would likewise be useful.

Spectroscopic and computational work by Tarczay et al. has examined the

chemistry of n-propylperoxy and isopropylperoxy radical, in each case finding multiple

conformers and noting the effect of those conformers on the A-X transitions of each

species.7 In tandem with their continuing cavity-ringdown spectroscopy (CRDS) work

on the subject, we have explored the conformations of n-propylperoxy radical and

potential effects these conformations could have on the unimolecular decomposition of n-

propylperoxy radical.8,9

In considering the overall chemistry of n-propylperoxy radical, it is important to

address its multiple conformations. n-Propylperoxy radical contains two dihedral angles

of particular interest, in this study: C−C−C−O and C−C−O−O. While multiple dihedrals are of course present in the molecule, these two particular angles afford rotation of the internal chain. Generally speaking, each dihedral angle accounts for three geometries.

Looking down a given bond, if we were speaking in typical organic terms, these would be the three staggered conformations; two of these are “gauche” geometries (60 degrees and −60 degrees) and the final is “trans” (180 degrees). To use a useful shorthand, a dihedral angle can either be clockwise gauche (+60 degrees, g), counterclockwise gauche

(-60 degrees, g’), or trans (180 degrees). Thus, with two dihedral angles, there would be

289

nine (32 = 9) potential conformers: tG, tG’, gT, g’T, gG’, g’G, gG, g’G’, tT. Each

rotamer is here designated first by rotation around the CCOO bond and second by

rotation around the CCCO bond. For instance, the gT conformer of n-propylperoxy

radical demonstrates a CCOO dihedral of +60 degrees and a CCCO dihedral of 180

degrees. Of these nine conformers, eight each will exist as one enantiomer of a pair; the

remaining all-trans conformer will be identical to its mirror image. Thus, there are five

likely, energetically-disparate conformers (Figure 6.2): tG, gT, gG’, gG, and tT. (We note here that methyl rotation, around the H−C−C−C dihedral angle, also occurs for this species; however, as all resulting conformers would be degenerate, we will ignore these to simplify further discussion.)

290

Figure 6.2. Conformeric possibilities for n-propylperoxy radical. The first letter refers to rotation around the C−C−O−O dihedral angle, and the second letter refers to rotation around the C−C−C−O dihedral angle. These angles can take one of three orientations: clockwise gauche (+60 degrees, denoted by g), counterclockwise gauche (−60 degrees, g’), or trans (180 degrees, t).

Interconversion between the rotamers of n-propylperoxy radical occurs via

transition states between the staggered conformations (i.e. through its eclipsed

conformations). Again, for each dihedral angle, three potential eclipsed forms exist. In

this case, our primary interest is locating the highest-energy eclipsed form for a given dihedral angle; the barrier to rotation around the bond in question can be quantified as the difference between the energies of the highest-energy eclipsed form and the lowest- energy staggered form. Typically, these barriers are not high; ethane demonstrates a

291

barrier height of 3.0 kcal/mol, while the central C−C bond of butane sees a rotational

barrier of 4.5 kcal/mol.10 Larger alkyl substituents lead to higher steric interactions and

higher barriers to rotation; heteroatoms often see competing trends between atomic size

and electronic effects.

We have generated the rotational profiles for the two key dihedral angles in n-

propylperoxy radical in multiple ways, and can compare the relevant rotational barriers to

other energies and enthalpies of reaction; the relative populations of the various

conformers can also be generated from these values. Additionally, HO2 extrusion

pathways available to both ethylperoxy and n-propylperoxy radical have been explored

via a variety of methods and basis sets, to determine the most useful computational

approach. In this work, we will discuss the conformations of n-propylperoxy radical and

the implications of its rotational barriers in the context of its overall reactivity.

6.2 Computational Methods

All geometry optimizations and vibrational frequency calculations were

completed with Gaussian 0311 at the Ohio Supercomputer Center. The B3LYP12,13 and mPW1K14 hybrid density functional theory (DFT)15 methods were used with the 6-

31+G** basis set for all geometry optimizations and vibrational frequency calculations;

the B3LYP method has been used successfully in a wide range of applications, but tends

to underpredict barrier heights, while the mPW1K method has been developed with a

292

functional optimized for kinetic applications. Additionally, CCSD(T)16 and CBS-QB3

single-point energy calculations17 were used to calibrate our hybrid DFT approach.

Vibrational frequencies were calculated for each stationary point to characterize

these structures as minima or transition states. The unscaled vibrational frequencies were

used to calculate the thermodynamic corrections to the enthalpy and free energy. Once

obtained, zero-point vibrational energy corrections were scaled by factors of 0.9806 and

0.9515, respectively, for B3LYP and mPW1K results.18 The enthalpy at each

temperature was determined from the single-point energy, the thermal correction to the

enthalpy, and the scaled zero-point energy, while the free energy at each temperature also

included the entropic correction to the free energy. Enthalpies at 0 K were calculated

from the single-point energy and the scaled zero-point energy.

Transition states were connected to reactants and products by either using

intrinsic reaction coordinate (IRC)19 searches or displacing the relevant geometries by ±

20% along the reaction coordinate characterized by the imaginary vibrational frequency,

then carefully optimizing (opt = calcfc) these resulting geometries to their corresponding

minima. For the doublet radical species and many transition states, spin contamination

() was negligible (no greater than 0.80). The methods employed in this work have

demonstrated an ability to minimize spin contamination for the radicals of interest,

keeping the computed values near the expected value of 0.75.

The Boltzmann distribution (Equation 6.2) was used to determine the relative

populations of the conformers of interest:

293

Δ− Bi TkE )/( ieg Ni = (6.2) Δ− Bj TkE )/( ∑ j eg j

Here, ΔEi is the energetic quantity of interest of conformer i, relative to the energy of the

global minimum structure (set to 0.0); gi is the structural degeneracy; kB is Boltzmann's

constant; T is temperature (298 K); and j runs over all unique conformers of the peroxy

radical of interest. Population distributions were determined referring to the energies,

enthalpies, and free energies.

All structures, vibrational frequencies, energies, thermal corrections to the

enthalpy and free energy, values, and rotational constants can be found in the

Supporting Information, for each geometry obtained in this study.

6.3 Results and Discussion

All possible transition state geometries (energetic maxima) were generated by

eclipsing each of three bonds for each dihedral rotation; correspondingly, all potential

minima were also identified. Rotational profiles for n-propylperoxy radical were

generated via mPW1K/6-31+G** and B3LYP/6-31+G** calculations (Figures 6.3 − 6.9),

both manually, using the independently-generated geometries, and computationally, via

scans of the relevant dihedral angles. The populations of each conformer were calculated

(Table 6.1), using their relative energies and the Boltzmann distribution. Potential HO2 losses (empirical evidence supports hydroperoxyl radical loss as a key step in the combustion of these species) were explored for both ethylperoxy and n-propylperoxy

294

radicals (Tables 6.1 and 6.2). Finally, this work was completed in tandem with a study of

the full decomposition of n-propylperoxy radical; some relevant energies are included for

reference (Figure 6.10).

6.3.1 Rotational profiles.

B3LYP/6-31+G** (Figure 6.3) and mPW1K/6-31+G** calculations (Figure 6.4) demonstrated that rotation around the C2-C3 bond of n-propylperoxy radical (via the

dihedral C−C−C−O) occurred with the most significant barrier height of the potential

rotations, between 4.5 − 5.0 kcal/mol.

gT gG gG’

Figure 6.3. Rotational profile for the C2−C3 bond of n-propylperoxy radical (rotating C−C−C−O dihedral while holding C−C−O−O dihedral constant at 60 degrees) calculated at the B3LYP/6-31+G** level of theory and reported as a function of electronic energy (kcal/mol).

295

Figure 6.4. Rotational profile for the C2−C3 bond of n-propylperoxy radical (rotating C−C−C−O dihedral while holding C−C−O−O dihedral constant at 60 degrees) calculated at the mPW1K/6-31+G** level of theory and reported as a function of electronic energy (kcal/mol).

As can be seen, rotation along the C−C bond results in formation of three minima: gT, gG’, and gG. The three intermediate transition states were isolated. Of these three, the transition state with the highest cost was intermediate between the gG’ and gG geometries (Figure 6.5). Here, the methyl group eclipses the peroxy moiety; the size of these two groups causes the greatest steric effect and thus highest rotational barrier. The

296

other two transition states were degenerate; in each, the eclipsing interactions were less, always involving at least one H atom.

CH O 3 O

H H H H

Figure 6.5. Visual representations of highest-energy rotational conformer for rotation around the C−C−C−O dihedral in n-propylperoxy radical.

Compared to C2−C3 bond rotation, rotation around the C3−O bond occurred with a low barrier height, as predicted by both hybrid DFT methods (Figures 6.6 and 6.7). In this case, the three energetic minima are the tT, g’T, and gT conformers of n-

propylperoxy radical. As with C−C bond rotation, three transition states were identified; two were at significantly lower energies than the third. The highest energy barrier occurred through a transition state in which the terminal oxygen atom eclipsed the rest of the carbon backbone, as pictured (Figure 6.8), but the other barriers were relatively small.

297

g’T gT tT

Figure 6.6. Rotational profile for the C3−O bond of n-propylperoxy radical (rotating C−C−O−O dihedral while holding C−C−C−O dihedral constant at 180 degrees), calculated at the B3LYP/6-31+G** level of theory and reported as a function of electronic energy (kcal/mol).

298

Figure 6.7. Rotational profile for the C3−O bond of n-propylperoxy radical (rotating C−C−O−O dihedral while holding C−C−C−O dihedral constant at 180 degrees), calculated at the mPW1K/6-31+G** level of theory and reported as a function of electronic energy (kcal/mol).

H LP

LP O CH CH H 2 3

Figure 6.8. Visual representations of highest-energy rotational conformer for rotation around the C−C−O−O dihedral in n-propylperoxy radical.

299

Finally, a profile of methyl rotation (Figure 6.9) was obtained, and validated our supposition that this rotation would yield three degenerate structures, which would not contribute to the conformeric distribution. As is typical for methyl rotation, this process occurred with a barrier height of 2.7 kcal/mol. The B3LYP profile is shown; the mPW1K information obtained and was identical, so is omitted.

Figure 6.9. Rotational profile for methyl rotation in n-propylperoxy radical, calculated at the B3LYP/6-31+G** level of theory and reported as a function of electronic energy (kcal/mol).

It is important to note that the rotations modeled here did not involve all the conformers as distinct energetic minima; rather, these calculations were run to examine the effect of dihedral rotation on a given starting conformer. Thus, the gT conformer of

300

n-propylperoxy radical exhibits a rotational barrier of roughly 5.0 kcal/mol around the

C2−C3 bond, while the g’T conformer exhibits a rotational barrier of 2.5 − 3.0 kcal/mol

around the C3−O bond. These numbers provide an illustration of common barrier heights

between rotamers; other work in our group examined the conformeric population

distribution, which took all five rotamers into account.

Overall, the rotational barriers did not provide any surprising results. Sterics played the major role in dictating rotational barriers; the process that would incur the

largest amount of strain was rotation around the C2−C3 bond, as it entailed eclipsing

interactions between a methyl group and the peroxy chain. Rotation around the C3−O

bond involved minor eclipsing interactions (often as small as that of a hydrogen-atom and

a lone pair); the least favorable form involved the eclipsing of the terminal oxygen atom

with the C2 atom, which did not involve any great steric interactions.

6.3.2 Conformer populations

Concurrent work in our group used CBS-QB3 calculations, along with B3LYP

and mPW1K calculations, to determine the overall distribution between the conformeric

forms of n-propylperoxy radical. Those energetics are reproduced here (Table 6.1).

301

n-Propylperoxy radical CBS-QB3 B3LYP/6-31+G** mPW1K/6-31+G**

Rotamer Degeneracy ΔG298 % ΔG298 % ΔG298 %

gG’ 2 0.41 14.0 0.62 11.8 0.57 11.7

gG 2 0.00 28.1 0.31 19.8 0.22 21.2

tG 2 0.04 26.4 0.11 27.8 0.06 28.0

gT 2 0.21 19.6 0.21 23.7 0.15 23.8

tT 1 0.10 11.9 0.00 16.8 0.00 15.4

Table 6.1. Boltzmann distributions for the five rotamers of n-propylperoxy radical, via their free energy differences (kcal/mol) and relative degeneracies at 298 K, as calculated at the CBS-QB3 and hybrid DFT levels of theory.

The alert reader may notice discrepancies in the global minima predicted by these calculations relative to the rotational profiles. This is attributed to the fact that each calculation relies on a different type of energy; the rotational profiles were generated based solely on the electronic energies (at 0 K) of the conformers, while these populations were generated from the free energies at 298 K. Thus, the two quantities are not immediately comparable. The population information is undoubtedly more quantitatively accurate; the rotational profiles were primarily of qualitative interest.

302

The experimental work of Zalyubovsky et al. did see multiple conformers of the

n-propylperoxy radical, validating these theoretical results. Three distinct peaks were

evident on the CRDS spectrum of interest; we hypothesize that these could be attributed

to: gG, tG, and tT (first peak), gT (second peak), and gG’ (third peak).

6.3.3 Competition between rotational barriers and reaction energetics

A key question in this study involved how the rotational barriers between

rotamers compared to the kinetic barriers of the reactions available to n-propylperoxy

radical. That is: would interconversion between reactant conformers play a major role in

affecting kinetics? Along with this conformational work, an exhaustive study of the

unimolecular decomposition of n-propylperoxy radical was undertaken by others in the group. In particular, the internal H-atom transfers of n-propylperoxy radical were

explored (Figure 6.10).

303

42.6 45.9 40.8

32.4 35.7 31.7 27.5 CH3CH2CH2 + O2 37.0 31.4 23.8 30.8 31.3 26.7 36.1 23.2

0.0 B3 LYP/ 6 - 3 1 + G* * 0.0 mPW1K/6-31+G** 0.0 CBS-QB3

Figure 6.10. Initial barrier heights for unimolecular decomposition of n-propylperoxy radical, expressed in terms of ΔH(298 K), kcal/mol, relative to n-propylperoxy radical. Figure courtesy of John K. Merle (J. Phys. Chem. A 2005, 109, 3637-3646). Reprinted with permission of J. Phys. Chem. A.

Clearly, the kinetic barriers to reaction were substantially higher than any internal barrier to rotation. Thus, at 298 K, rotamer interconversion is expected to be negligible in considering the overall chemistry of n-propylperoxy radical. If these results were extended to a prediction of rate coefficients for the processes of n-propylperoxy radical, it still would be valuable to consider the various rotamers afforded to the reactant and transition state, as noted in the work of Vereecken and Peeters; however, this study was

304

primarily concerned with an overall description of n-propylperoxy radical’s unimolecular

decomposition, rather than the kinetic rates of those processes.

The computational study of n-propylperoxy radical’s unimolecular decomposition

found that the likeliest pathway involved 1,4-H-atom transfer concomitant with loss of

hydroperoxyl radical; this was shown to occur with a barrier height of 30.8 kcal/mol via

CBS-QB3 calculations (27.5 kcal/mol at the B3LYP/6-31+G** level). These findings

demonstrated the comparability of hybrid DFT results to high-level calculations and, moreover, were in agreement with experimental work on the radical; DeSain et al. have monitored the decomposition of n-propylperoxy radical by studying the production of hydroperoxyl radical and proposed an activation barrier of 26.0 kcal/mol for the

20 elimination/HO2 production reaction.

6.3.4 Computational approaches

One final aspect of this work involved the examination of several DFT and

higher-level calculations in predicting the chemistry of ethylperoxy and n-propylperoxy

radicals. An ongoing challenge for computational chemists involves the balance of

quantitative accuracy with computational resources. With the small species of interest in

this work, many calculations could be run to determine a useful method and basis set for

use in further study of small alkylperoxy radicals.

In particular, as hydroperoxyl radical is a commonly-observed product of both

ethylperoxy and propylperoxy radical reactions, potential steps involving generation of

this radical were modeled for both species (Tables 6.2 and 6.3).

305

O O C C O C + O O C C C O

‡ o Method/Basis Set ΔH 298K (kcal/mol) ΔH298K (kcal/mol) B3LYP/6-31G* a 12.2 2.50 b B3LYP/6-311+G** 10.6 −0.41 b B3LYP/aug-cc-pVTZ 11.0 −0.64 mPW1K a 18.9 6.76 CCSD(T)/aug-cc-pVDZ b,d 18.8 3.65 c,d CCSD(T)/aug-cc-pVDZ 19.6 −1.43

‡ o Method/Basis Set ΔG 298K (kcal/mol) ΔG 298K (kcal/mol) a B3LYP/6-31G* 12.5 −7.22 b B3LYP/6-311+G** 10.9 −10.1 b B3LYP/aug-cc-pVTZ 11.3 −10.4 a mPW1K 19.1 −3.19 b,d CCSD(T)/aug-cc-pVDZ 19.1 −6.08 c,d CCSD(T)/aug-cc-pVDZ 19.9 −11.2

‡ ‡ Method/Basis Set ΔH0K = ΔG 0K (kcal/mol) ΔH0K = ΔG0K (kcal/mol) B3LYP/6-31G* a 12.9 3.77 B3LYP/6-311+G** b 11.3 0.86 B3LYP/aug-cc-pVTZ b 11.7 0.63 mPW1K a 18.9 6.80 CCSD(T)/aug-cc-pVDZ b,d 19.5 4.92 c,d CCSD(T)/aug-cc-pVDZ 20.3 −0.16 a Fully optimized geometry b Single-point energy calculation using B3LYP/6-31G* geometry c Single-point energy calculation using mPW1K geometry d Previous research with CCSD gives ΔH‡= 15.1 kcal/mol, using different geometry optimization [CCSD(T)/TZ2P//CCSD(T)/DZP]. (Schaefer et al. J. Phys. Chem. A 2000, 104, 9823-9840)

Table 6.2. Enthalpies and energies of activation and reaction (kcal/mol) for the scission of ethylperoxy radical into ethene and hydroperoxyl radical, calculated via several methods and basis sets.

306

-1 υ1 = 915i cm (B3LYP) -1 υ1 = 1061i cm (mPW1K)

O O O C 2 4 C C + C 5 3 1 O C C

‡ ‡ o o Method/Basis Set ΔH 298K ΔG 298K ΔH298K ΔG298 K (kcal/mol) (kcal/mol) (kcal/mol) (kcal/mol)

B3LYP/6-31G* a 35.1 35.8 6.94 -1.25 B3LYP/6-311+G** b 33.1 33.8 6.11 -2.09 B3LYP/aug-cc-pVTZ b 33.3 34 5.75 -2.45 mPW1K/6-31+G** a 41.2 41 7.61 -1.98 a Fully optimized geometry b Single-point energy calculation using B3LYP/6-31G* geometry

Table 6.3. Enthalpies and energies of activation and reaction (kcal/mol) for the cyclization reaction of n-propylperoxy radical, dissociating into cyclopropane and hydroperoxyl radical, calculated via several methods and basis sets.

A variety of single-point energy calculations were completed, using the B3LYP

and mPW1K geometry optimizations. These values varied to a wide extent; generally,

the predicted barrier heights increased with the size of the basis set, while the overall

thermodynamics of the reaction became more exothermic and exoergic. In the case of

the ethylperoxy radical reaction, previous research values were available; Schaefer et al.

have reported a barrier height of 15.1 kcal/mol for the simultaneous 1,4-H-atom transfer and HO2 elimination, which they have supported with experimental evidence as the

307

likeliest pathway for ethylperoxy radical degradation. This value actually falls in the

middle range of the energies we calculated; their CCSD energy was calculated with a

different energy optimization, which probably accounts for the difference. The B3LYP

calculations did provide reasonable values at significantly lesser energetic costs than the

CCSD(T) calculations, and concurrent work in our group demonstrated that it also

provided quantitatively reasonable values for the overall unimolecular decomposition of

n-propylperoxy radical via 1,4-H-atom transfer concurrent with HO2 elimination, which was deemed to be the most likely reaction pathway. Thus, the hybrid DFT methods

B3LYP and mPW1K show promise for modeling intramolecular reactions of alkylperoxy radicals, while B3LYP is often said to underestimate barrier heights in intermolecular reactions, it is much improved in considering barrier heights of intramolecular processes.

6.4 Conclusions

Using the hybrid DFT methods B3LYP/6-31+G** and mPW1K/6-31+G**, five

main conformers were identified as contributing to the overall chemistry of n-

propylperoxy radical: gG’, gG, tG, gT, and tT, which were all doubly degenerate, except

for the all-trans conformer. These rotamers were shown to rapidly interconvert, with

rotational barriers no higher than 5.0 kcal/mol in any case, via rotation of the C2−C3 and

C3−O bonds. Methyl rotation could also occur but did not affect the overall number of rotamers. Rotation around the C2−C3 bond saw the highest rotational barrier due to the

greatest number of steric interactions. Due to the rapid interconversion process, all five

rotamers would contribute to the overall population at 298 K. This was verified with

308

concurrent experimental findings. If this work were to be extended to kinetic

calculations, these rotamers would also play a significant role in calculating the rate

coefficient for n-propylperoxy radical, as explained by Vereecken and Peeters for the

analogous structure of 1-butoxy radical.

The internal H-atom shifts available to n-propylperoxy radical, as well as its resulting unimolecular decomposition, have also been calculated in tandem with this

work; 1,5-H-atom transfer occurred with the lowest barrier due to its ability to occur via a

six-membered transition state, while 1,4-H-atom transfer concurrent with hydroperoxyl

radical elimination ultimately contributed to the decomposition of n-propylperoxy radical

to the greatest extent. For our purposes, what is most notable is that all reaction steps

occurred with much higher barrier heights than did the rotamer interconversions.

309

References for Chapter 6

1 (a) Wallington, T. J.; Dagaut, R.; Kurylo, M. J. Chem. Rev. 1992, 92, 667-710; (b)

Wallington, T. J.; Nielsen, O. J.; Sehested, J. “Reactions of Organic Peroxy Radicals in

the Gas Phase,” in Peroxyl Radicals, Ed. Alfassi, Z. John Wiley and Sons: New York,

1997.

2 Ignatyev, I. S.; Xie, Y.; Allen, W. D.; Schaefer, H. F., III. J. Phys. Chem. A 1997, 107,

141.

3 Rienstra-Kiracofe, J. C.; Allen, W. D.; Scaefer, H. F., III. J. Phys. Chem. A 2000, 104,

9823-9840.

4 (a) Wagner, A. F.; Slagle, I. R.; Sarzynski, D.; Gutman, D. J. Phys. Chem. 1990, 94,

1853; (b) Kaiser, E. W.; Wallington, T. J.; Andino, J. M. Chem. Phys. Lett. 1990, 168,

309; (c) Kaiser, E. W.; Rimai, L.; Wallington, T. J. J. Phys. Chem. 1989, 93, 4094; (d)

Kaiser, E. W.; Lorkovic, I. M.; Wallington, T. J. J. Phys. Chem. 1990, 94, 3352; (e)

Dobis, O.; Benson, S. W. J. Am. Chem. Soc. 1993, 115, 8798; (f) Kaiser, E. W. J. Phys.

Chem. A 2002, 106, 1256; (g) Kaiser, E. W. J. Phys. Chem. 1995, 99, 707; (h) Clifford,

E. P.; Farrell, J. T.; DeSain, J. D.; Taatjes, C. A. J. Phys. Chem. A 2000, 104, 11549.

5 (a) Orlando, J. J.; Tyndall, G. S.; Bilde, M.; Ferronato, C.; Wallington, T. J.; Vereecken,

L.; Peeters, J. J. Phys. Chem. A 1998, 102, 8116; (b) Vereecken, L.; Peeters, J. J. Phys.

Chem. A 1999, 103, 1768; (c) Vereecken, L.; Peeters, J.; Orlando, J. J.; Tyndall, G. S.;

Ferronato, C. J. Phys. Chem. A 1999, 103, 4693.

6 Vereecken, L.; Peeters, J. J. Chem. Phys. 2003, 119, 5159-5170.

7 Tarczay, G.; Zalyubovsky, S. J.; Miller, T.A. Chem. Phys. Lett. 2005, 406, 81-89.

310

8 Zalyubovsky, S.; Glover, B.; Miller, T.; Hayes, C.; Merle, J.; Hadad, C.; J. Phys. Chem.

A, 2005, 107, 1308-1315.

9 Merle, J. K.; Hayes, C. J.; Zalyubovsky, S. J.; Glover, B.; Miller, T. A.; Hadad, C. M. J.

Phys. Chem. A. 2005, 109, 3637-3646.

10 Eliel, E.; Wilen, S. H.; Doyle, M. P. Basic Organic Stereochemistry. New York :

Wiley-Interscience, 2001.

11 Gaussian 03, Revision C.02, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G.

E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.;

Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi,

M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.;

Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.;

Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;

Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;

Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.;

Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain,

M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.;

Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.;

Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham,

M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.;

Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A. Gaussian, Inc., Wallingford CT,

2004.

311

12 (a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648; (b) Becke, A. D. Phys. Rev. A. 1988,

38, 3098-3100.

13 Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B 1998, 37, 785-789.

14 Lynch, B. J.; Fast, P. L.; Harris, M.; Truhlar, D. G. J. Phys. Chem. A 2000, 104, 4811.

15 (a) Parr, R. G.; Yang, W. Density Functional Theory in Atoms and Molecules; Oxford

University Press: New York, 1989. (b) Labanowski, J. W.; Andzelm, J. Density

Functional Methods in Chemistry; Springer: New York, 1991.

16 Cizek, J. J. Chem. Phys. 1966, 45, 4256.

17 (a) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A., Jr. J. Chem. Phys. 1996,

104, 2598-2619; (b) Petersson, G. A.; Frisch, M. J. J. Phys. Chem. A 2000, 104, 2183-

2190.

18 Scott, A. P.; Radom, L. J. Phys. Chem. 1996, 100, 16502-16513.

19 (a) Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1989, 90, 2154. (b) Gonzalez, C.;

Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523.

20 DeSain, J.; Klippenstein, S.; Miller, J.; Taatjes, C. J. Phys. Chem. A 2001, 105, 4415-

4425.

312

CHAPTER 7

THE UNIMOLECULAR DECOMPOSITION OF PROPAN-1-OL-1-PEROXY RADICAL: EFFECTS OF FUNCTIONALIZATION ON PEROXY RADICAL PATHWAYS AND IMPLICATIONS FOR BIOFUEL COMBUSTION

7.1. Introduction

Biofuels comprise a class of alternative fuels that have promise for decreasing

society’s dependence on fossil fuels and minimizing harmful emissions. At the same

time, the increased size and functionalization of these species present challenges to combustion chemists. Comprehensive mechanisms of a given fuel play a vital role in optimizing engine performance and fuel blends. Historically, mechanism development is a collaborative process; for instance, the mechanisms of the primary reference fuels1 used to model gasoline combustion rely heavily on pathways for methane2 that have been

extant for decades. While mechanisms available for hydrocarbon fuels have been

extensively compiled, the and esters of interest to biofuel chemistry3 have only

recently begun to undergo similar mechanistic analysis.

Combustion mechanisms for some smaller oxygenated compounds have been

developed; the most common targets of interest have been alcohols and esters. Marinov

has developed an exhaustive mechanism for ethanol’s oxidation.4 Cavalli et al. have

used FTIR spectroscopy to examine the initial reaction of hydroxyl (HO•) radical with 1-

313 butanol.5 Recent experimental work by McEnally and Pfefferle has proposed that

butanol combustion primarily occurs via a complex fission reaction rather than H-atom

abstraction.6 Curran et al.7 and Gail et al.8 have compiled high-temperature combustion data for methyl butanoate; Good and Francisco have modeled the tropospheric oxidation mechanism of methyl formate.9 Metcalfe et al. used high-level calculations to explore

10 the combustion of C5H10O2 esters. While the majority of these studies were interested

in oxygenated compounds as part of useful fuel blends, the more recent studies have

shifted towards an interest in these species as fuels in their own right. Several reviews

are available on more general concepts of biomass fuels. Schwab et al. has discussed vegetable oils as potential biodiesel sources;11 the general fuel properties of biomass

compounds have been compiled by Graboski and McCormick,12 Srivastava and Prasad,13 and Lin et al.14 Ma and Hanna recently reviewed the general benefits of biofuel

chemistry.15

Like all fuel molecules, biofuels would have a propensity for oxidation via a

peroxy radical intermediate at lower combustion temperatures; moreover, radicals formed

via these fuels could undergo comparable reactions in the atmosphere. Peroxy radicals

would thus play a significant role in the performance and decomposition of these

important fuel species. The basic steps of peroxy radical chemistry have been well-

characterized (Figure 7.1).16,17

314 (3) A+ROOH RO + HO AH (2) (1) O R+O2 RO2 CH +HO (4) 2 (5) R' R'CH OOH 2 (6)

R(OO)CH2OOH

(7)

R(OOH)CHO + HO

(8)

R(O)CHO + HO

Figure 7.1. Common steps of peroxy radical chemistry.

A given radical can add oxygen to form a peroxy radical (1). At low temperature and with collisional stabilization, this reactive species has several options, most of which begin with a hydrogen-atom transfer. This can be an intermolecular process involving reaction with another molecule (2), yielding a hydroperoxide and another alkyl radical; the hydroperoxide can dissociate into alkoxy and hydroxyl radicals via a chain-branching step (3). If the alkyl chain of the original peroxy radical is sufficiently long, the hydrogen-atom transfer can also occur intramolecularly (4), forming a hydroperoxyl alkyl radical; this latter species can decompose unimolecularly via cyclizations (5) or β- scissions; it can also add O2 at the incipient radical center (6) to yield a new peroxy

315 radical that ultimately gives functionalized products (7, 8). As seen even in this cursory

overview, the steps available to the hydroperoxyalkyl radical result in chain-branching

and have been implicated in autoignition processes.18 The specific nature of the final

products depends on the identity of the parent fuel molecule: as biofuels are oxygenated species, their peroxy radicals would have the potential for even more diverse pathways.

Besides these self-reactions, peroxy radicals can react with a variety of other species, in low-temperature combustion environments. This is especially evident in atmospheric chemistry. They can react with NO to form oxyradicals and NO2; NO2 can

then participate in ozone production.19 Ozone overproduction thus becomes a concern

when hydrocarbon emissions are high. Other reactions with atmospheric implications

• involve reactions with hydroperoxyl (HO2 ) radical and other peroxy radicals to form alcohols, aldehydes, ketones, and hydroperoxides. Again, the presence of functional groups, such as those present in biofuels, could have an impact on known atmospheric processes, as well as introducing qualitatively different products.

The hydroperoxyalkyl radicals are generally referred to via the convention

QOOH, which will be adopted in the subsequent discussion of our own work. For a given peroxy radical, several isomerized products QOOH can be identified; these are distinguished by a label (1,xn), where x denotes the relative position of the new radical center and n represents its type (primary, secondary, etc.). For instance, Q(1,5p)OOH indicates a 1,5-H atom transfer, resulting in a primary radical at position 5 relative to the initial radical center.

316 Ethylperoxy radical is the smallest system capable of undergoing an internal H-

atom transfer of any significance; n-propylperoxy radical is the smallest system wherein

this internal H-atom transfer can be accomplished via a six-membered transition state.

Both of these systems are valuable as models for larger hydrocarbon systems. Ethyl

20 radical + O2 has been studied experimentally and theoretically; Ignatyev et al.

demonstrated that its most likely decomposition would yield ethene and hydroperoxyl

radical via a concerted mechanism.21 With n-propylperoxy radical, the 1,4- isomerization, 1,4-elimination, and 1,5-isomerization pathways have all been studied.22

The production of hydroperoxyl23 and hydroxyl radicals24 have been modeled, as well.

Again, a concerted H-atom transfer concurrent with elimination was shown to be the dominant mechanism of unimolecular decomposition; propene and hydroperoxyl radicals were consistently the most common empirically observed products.25

Similar reactions have been studied in larger systems; however, these have all

tended to be hydrocarbons: fewer studies have examined how the peroxy radical chemistry of a system might be tempered or promoted by the presence of any additional functional groups, such as those found in biomass compounds. McEnally and Pfefferle showed that butanol decomposition proceeds primarily through internal reactions.

Metcalfe et al. saw that methyl butanoate and ethyl propanoate were both prone to intramolecular scissions as primary routes of decomposition.10 Thus, while few relevant

studies of oxidative decomposition have been completed, the extant work supports unimolecular decompositions as key processes in understanding biofuel chemistry.

317 We are interested in further exploring peroxy radical chemistry as it pertains to

the oxidative decomposition of functionalized compounds. Our group has completed

studies on various peroxy radicals formed by alkyl, aromatic, and heterocyclic parent

compounds. Merle et al. have modeled the rearrangement pathways available to n-

propylperoxy radical.26 In terms of initial reactivity of n-propylperoxy radical, five

pathways were shown to be feasible at low temperatures: dissociation back to reactants,

• 1,3-H transfer, 1,4-H transfer, 1,4-H transfer concurrent with HO2 elimination, and 1,5-H

transfer. The resultant decomposition steps available to each QOOH transfer product

were also explored. 1,3-H transfer directly yielded combustion products propanal and

• • HO , and 1,4-H transfer/HO2 elimination yielded propene and hydroperoxyl radical. For

the isolable H-atom transfer radicals, several options were available. The Q(1,4s)OOH

• • isomer could cyclize to yield cyclopropane and HO2 , dissociate into propene and HO2 , or cyclize to methyloxirane and HO•. The Q(1,5p)OOH isomer could cyclize to oxetane and HO•; dissociate into ethene, formaldehyde, and HO•; or undergo a subsequent H-

atom transfer to yield the Q(1,4s)OOH radical.

Overall, it was seen that 1,5-H transfer experienced the lowest kinetic barrier, due

to the six-membered transition state through which it could achieve its transfer; however,

subsequent barrier heights in each set of decomposition pathways made the 1,4-H-

• transfer/HO2 elimination pathway the likeliest overall decomposition. This finding was

in agreement with experimental studies that have identified propene and hydroperoxyl

27 radical as the major products of propyl radical + O2; generally, the products achieved via 1,3- and 1,5-H atom transfers, such as cyclic and hydroxyl radical, are less

318 commonly observed. We have also shown that multiple conformers are possible for this

reactive intermediate and would contribute to its overall chemistry at 298 K, helping to

validate a collaborative cavity-ringdown spectroscopy experiment by Zalyubovsky et

al.28 This was in keeping with the previous kinetic analysis of the 1,5-H atom shift in 1-

butoxy radical, completed by Vereecken and Peters,29 who noted the necessity of

incorporating all conformers to obtain a valid rate coefficient.

To complete this work, we used both density functional theory30 (B3LYP31,32 and

mPW1K33) and complete basis set34 (CBS-QB3) methods; B3LYP/6-31+G** was shown

to replicate nearly all qualitative trends predicted by CBS-QB3, and in many cases was quantitatively comparable as well, while mPW1K showed a tendency to overpredict barrier heights.26

We now seek to explore the implications of our findings for functionalized

compounds. The hydroxyl and ester moieties are two common functional groups found

in biomass compounds; in this initial study, we will approach the simplest extension of

our initial work on n-propylperoxy radical, examining the combustion of 1-propanol

rather than propane. We will determine the qualitative changes seen in product

distribution when the fuel of interest is an alcohol, rather than a simple alkane. We also

are interested in the quantitative effect this additional functionality can have on relative

kinetic barriers and reaction thermochemistry. Many concerns are duplicated between

studies. The resultant peroxy radical of 1-propanol will exist in several conformeric

forms. As with our previous study of n-propylperoxy radical, we will calibrate our

density functional theory approach, using methods such as B3LYP and mPW1K, against

319 complete basis set methods (CBS-QB3), in order to suggest a useful and computationally

economical approach to the chemistry of similar species.

7.2. Computational Methods.

All geometry optimizations and vibrational frequency calculations were

completed with Gaussian 0335 at the Ohio Supercomputer Center. The B3LYP32,32 and

mPW1K hybrid density functional theory (DFT) methods were used with the 6-31+G**

basis set for all geometry optimizations and vibrational frequency calculations; the

B3LYP method has been used successfully in a wide range of applications, but tends to

underpredict barrier heights, while the mPW1K method has been developed with a

functional optimized for hydrogen-atom abstractions. Additionally, the Complete Basis

Set (CBS-QB3) method was used to generate overall potential energy surfaces for the

oxidation of 1-propanol and its subsequent peroxy radical pathways. This comprehensive

method uses a B3LYP geometry in conjunction with single-point energies from CCSD(T)

and complete basis set extrapolation approaches using Møller-Plesset perturbation theory,

in order to generate a CCSD(T) energy near the complete basis set level.

Vibrational frequencies were calculated for each stationary point to characterize

these structures as minima or transition states. The unscaled vibrational frequencies were

used to calculate the thermodynamic corrections to the enthalpy and free energy. Once

obtained, zero-point vibrational energy corrections were scaled by factors of 0.9806 and

0.9515, respectively, for the B3LYP and mPW1K results.36 The overall enthalpy at each

temperature was determined from the single-point energy, the thermal correction to the

320 enthalpy, and the scaled zero-point vibrational energy (ZPE) correction, while the overall

free energy at each temperature also included the entropic correction to the free energy.

Enthalpies at 0 K were calculated from the single-point energy and the scaled ZPE.

Transition states were connected to reactants and products by either using

intrinsic reaction coordinate (IRC)37 searches or displacing the relevant geometries by

±20% along the reaction coordinate characterized by the imaginary vibrational frequency, then carefully optimizing (opt = calcfc) these resulting geometries to their corresponding minima. For the doublet radical species and many transition states, spin contamination

() was negligible (no greater than 0.80). The methods employed in this work have

demonstrated an ability to minimize spin contamination for the radicals of interest,

keeping the computed values near the expected value of 0.75.

The Boltzmann distribution was used to determine the relative populations of the

conformers of interest:

Δ− Bi TkE )/( i eg Ni = Δ− Bj TkE )/( ∑ j eg j

Here, ΔEi is the energetic quantity of interest of conformer i, relative to the energy of the

global minimum structure (set to 0.0); gi is the structural degeneracy; kB is Boltzmann's

constant; T is temperature (298 K); and j runs over all unique conformers of the peroxy

radical of interest. Population distributions were determined referring to the energies,

enthalpies, and free energies.

321 7.3. Results and Discussion.

In propanol, three alkyl carbon sites are possible for initial hydrogen-atom loss

(Table 7.1). It has been well-established that the most reactive site in an alcohol is at the carbon directly adjacent to the hydroxyl group, and this was confirmed via our initial calculations, via both CBS-QB3 and DFT methods. Correspondingly, propan-1-ol-1- peroxy radical (Figure 7.2) is the likeliest low-temperature peroxy radical to be formed from the subsequent addition of O2 to the carbon-centered radical.

2 OH

3 1

H-atom abstraction site CBS-QB3 B3LYP mPW1K Carbon 1 95.5 92.9 91.7 Carbon 2 100.3 97.4 96.1 Carbon 3 101.7 100.1 98.2

Table7.1. Bond dissociation enthalpies (298 K, kcal/mol, CBS-QB3, B3LYP/6-31+G**, and mPW1K/ 6-31+G**) for each unique C─H bond in 1-propanol.

322 OH

O O

Figure 7.2. Propan-1-ol-1-peroxy radical.

7.3.1. Conformations. Propan-1-ol-1-peroxy radical contains three main dihedral angles of interest: C−C−C−Operoxy and C−C−O−Operoxy, involving the carbon skeleton and the

38 peroxy group; and C−C−C−Ohydroxyl, incorporating the OH moiety. For each of these

three dihedral angles, three distinct conformers are likely to be energetic minima: a

gauche orientation involving a dihedral angle of 60o, a gauche orientation of –60o, and a trans orientation of 180o. The various combinations of these dihedral angles are summarized in Table 7.2. In referring to these conformers, we will adopt a naming convention for the various dihedral angles, using a three letter suffix wherein the

C−C−C−Operoxy angle is represented by the first letter; C−C−O−Operoxy, by the second;

o and C−C−C−Ohydroxyl, by the third. The letter g will refer to a gauche angle of +60 , while g’ will refer to a gauche angle of –60o, and t to a trans angle of 180o. The second

of these three letters will be capitalized for easier comprehension. For instance, 1-

peroxy-tTt will be used to refer to the all-trans conformer of propan-1-ol-1-peroxy radical in which all three of the dihedral angles of interest are ~180o.

323 In total, a species with three dihedral angles can potentially generate 27 unique conformers; all of these except the all-trans form are present as enantiomers, such that there are 14 unique conformations available to the propan-1-ol-1-peroxy radical. These conformations could have implications in the spectroscopic analyses of biofuel-like compounds; moreover, as alluded to in the Introduction, Vereckeen has noted the implications of multiple conformers in providing accurate kinetic data.29 Species with all

27 of the possible dihedral combinations were generated in providing the starting geometries for our conformational analysis; these were all subsequently minimized and duplicate geometries were eliminated, using both CBS-QB3 and B3LYP/6-31+G** calculations. It was seen that five of these conformations (Table 7.3) were responsible for roughly 90% of the peroxy radical distribution: g’Gg’, gGt, gTg, tG’g’, and tGg’.

324

Conformer C−C−C−Operoxy (°) C−C−O−Operoxy (°) C−C−C−Ohydroxy (°) g'G'g' −60 −60 −60 g'G'g −60 −60 60 g'G't −60 −60 180 g'Gg' −60 60 −60 g'Gg −60 60 60 g'Gt −60 60 180 g'Tg' −60 180 −60 g'Tg −60 180 60 g'Tt −60 180 180 GG'g' 60 −60 −60 GG'g 60 −60 60 gG't 60 −60 180 gGg' 60 60 −60 gGg 60 60 60 gGt 60 60 180 gTg' 60 180 −60 gTg 60 180 60 gTt 60 180 180 TG'g' 180 −60 −60 TG'g 180 −60 60 TG't 180 −60 180 tGg' 180 60 −60 tGg 180 60 60 tGt 180 60 180 tTg' 180 180 −60 tTg 180 180 60 tTt 180 180 180

Table 7.2. Possible dihedral combinations considered in generating starting conformations of propan-1-ol-1-peroxy radical Three dihedral angles could each attain gauche (clockwise), gauche (counter-clockwise), or trans orientations, and the corresponding angles would be 60°, −60°, or 180°, respectively. Following minimization via CBS-QB3 and B3LYP/6-31+G**, and elimination of duplicate geometries, the major rotamers were identified.

325

Major Rotamer CBS-QB3 (%) B3LYP/6-31+G** (%) gGt 2.6 1.3 g’Gg’ 6.3 5.2 tG’g’ 21.1 18.1 gTg 27.0 26.1 tGg’ 35.4 35.7 Total 92.4% 86.4%

Table 7.3. Populations of the five most common rotamers for propan-1-ol-1-peroxy radical, using H0K data calculated via CBS-QB3 and B3LYP/6-31+G** energies and Boltzmann weighting factors. [The populations of the other rotamers are small (< 1%.)]

Both the CBS-QB3 method and B3LYP/6-31+G** identified the 1-peroxy-tGg’ conformer as the global minimum, and both methods predicted quantitatively similar amounts of the dominant five conformations (Figure 7.3).

326

1peroxy-g'Gg' 1peroxy-gGt 1peroxy-gTg

1peroxy-tG'g' 1peroxy-tGg'

Figure 7.3. Likeliest conformers of propan-1-ol-1-peroxy radical. The three-letter suffix denotes the rotation around the C−C−C−Operoxy, the C−C−O−Operoxy, and C−C−C−Ohydroxyl, respectively: g = gauche, clockwise (+60°), g’ = gauche, counter- clockwise (−60°), and t = trans (180°).

These conformers were all within 5 kcal/mol of one another; it was shown that interconversion would not have a significant effect on the reaction energetics of propan-

1-ol-1-peroxy radical.

7.3.2. Initial H-atom transfers. The five pathways described in Figure 7.4 are those most immediately available to propan-1-ol-1-peroxy radical, building from our group’s previous work on n-propylperoxy radical.

327

OH

+O2

OH 1,5-H transfer 1,3-H transfer OH OH

O O O OH OH O

1,4 H-transfer/ HO . elimination 2 1,4-H transfer

OH

+HO2

O OH

Figure 7.4. Isomerization pathways available to propan-1-ol-1-peroxy radical, based on previous work with n-propylperoxy radical.

Unless otherwise noted, we will discuss these routes in terms of their CBS-QB3 ΔH0K values, relative to the minimum conformer of propan-1-ol-1-peroxy radical, although a variety of energetic data are compiled in Tables 7.4─7.6. Figure 7.5 shows the initial kinetic barriers for the isomerization pathways of propan-1-ol-1-peroxy radical.

328

CBS-QB3 ΔH (0 K) ΔH (298 K) ΔG (298 K) Peroxy propanol (tGg’ conformer) 0.0 0.0 0.0 Peroxy propanol, 1,3-H reactant conformer 2.3 2.3 2.3 Peroxy propanol, 1,4-H reactant conformer 0.9 0.9 2.0 Peroxy propanol, 1,5-H reactant conformer 0.2 0.2 0.2 TS_1,3─H-atom transfer 45.7 41.6 42.0 TS_1,4─H-atom transfer 36.0 37.7 32.9 TS_1,4─elimination 36.3 32.4 32.5 TS_1,5─H-atom transfer 29.3 24.9 26.6 Propanoic acid + OH (complex) -39.9 -40.9 -43.1 Propanoic acid + OH (product) -34.1 -36.7 -47.1

Propenol + HO2 (complex) 12.5 25.3 13.4

Propenol + HO2 (product) 4.3 16.7 4.4

Q(1,4s)OOH 16.8 15.4 14.2 Cis-oxirane_reactant 15.4 14.0 12.8 TS (Q(1,4s)OOH → cis-oxirane + OH 26.6 23.7 22.7 Cis-oxirane_complex -5.7 -7.3 -8.2 Cis-1-hydroxy-2-methyl oxirane -3.0 -0.9 -10.4 Trans-oxirane_reactant 15.4 13.9 12.7 TS (Q(1,4s)OOH → trans-oxirane + OH 26.8 23.9 22.6 Trans-oxirane + OH (complex) -5.9 -7.6 -8.5 Trans-oxirane + OH 1.7 -1.3 -10.7 Propenol_reactant 14.9 13.5 12.3 TS (Q(1,4s)OOH → propenol + OH) 33.2 30.9 30.3

Propenol + HO2 (complex) 19.8 18.9 16.1

Propenol + HO2 (separated) 27.9 25.3 13.4

Propanal + HO2 19.7 16.7 4.4

Table 7.4. Enthalpies and free energies of reaction (kcal/mol) at 0 K and 298 K for the unimolecular decomposition of propan-1-ol-1-peroxy radical, calculated at the CBS-QB3 level of theory. See Figure 7.4 for the corresponding pathways. Table 7.4 continued on page 325.

329

Table 7.4, continued.

CBS-QB3 ΔH (0 K) ΔH (298 K) ΔG (298 K) H_transfer_reactant 0.1 0.2 0.2 TS (Q(1,5p)OOH → Q(1,4s)OOH) 52.3 56.8 56.6 Product conformer, Q(1,4s)OOH 15.0 17.1 15.5 Ethene_HCOOH_OH_reactant 14.9 16.9 16.1 TS (Q(1,5p)OOH → ethene, HCOOH, OH) 34.6 38.9 37.5 Ethene_HCOOH_OH_complex -20.7 -16.2 -23.2 Ethene + HCOOH + OH -18.2 -11.1 -32.0 Cyclopropanol_OH_reactant 14.7 16.6 16.1 TS (Q(1,5p)OOH → cyclopropanol +OH) 57.6 60.4 60.5 Cyclopropanol + OH 33.6 36.6 25.6 Oxetanyl_OH_reactant 13.2 14.7 14.6 TS (Q(1,5p)OOH → oxetane + OH) 32.9 35.3 35.5 Oxetanyl_OH complex -7.7 -7.0 -7.9 2-Hydroxy oxetane + OH -2.2 -1.6 -10.6

mPW1K/6-31+G** ΔH (0 K) ΔH (298 K) ΔG (298 K) Peroxy propanol (tGg’ conformer) 0.0 0.0 0.0 Peroxy propanol, 1,3-H reactant conformer 2.5 2.4 0.0(1) Peroxy propanol, 1,4-H reactant conformer 1.2 1.2 1.4 Peroxy propanol, 1,5-H reactant conformer 0.2 0.2 0.1 TS_1,3─H-atom transfer 50.7 46.6 46.9 TS_1,4─H-atom transfer 40.1 35.8 36.9 TS_1,4─elimination 39.3 34.1 34.0 TS_1,5─H-atom transfer 32.8 28.3 29.9 Propanoic acid + OH (Complex) -40.0 -41.4 -46.1 Propanoic acid + OH (product) -33.9 -36.8 -47.3

Propenol + HO2 (complex) 29.9 26.1 13.8

Propenol + HO2 (product) 21.3 17.1 4.3

Table 7.5. Enthalpies and free energies of reaction (kcal/mol) calculated at the mPW1K/6-31+G** level of theory at 0 K and 298 K for the unimolecular decomposition of propan-1-ol-1-peroxy radical. See Figure 7.4 for the corresponding pathways. Continued on page 326.

330

Table 7.5. continued.

mPW1K/6-31+G** ΔH (0 K) ΔH (298 K) ΔG (298 K) Q(1,4s)OOH 18.4 17.0 15.8 Cis_oxirane_reactant 17.1 15.9 14.5 TS (Q(1,4s)OOH → cis-oxirane + OH 31.3 28.3 26.9 Cis_oxirane_complex -7.5 -9.1 -10.5 Cis-1-hydroxy-2-methyl oxirane 0.8 -2.3 -12.1 Trans_oxirane_reactant 20.8 19.5 18.2 TS (Q(1,4s)OOH → trans-oxirane + OH 34.6 31.6 30.6 Trans-oxirane + OH (complex) -4.0 -5.7 -7.0 Trans_oxirane + OH 3.8 0.8 -8.8 Propenol_reactant 16.7 14.3 12.7 TS (Q(1,4s)OOH → propenol + OH) 37.6 34.0 32.9

Propenol + HO2 22.3 20.0 16.0

Propanal + HO2 29.9 26.1 13.8

Q(1,5p)OOH 21.9 0.2 0.1 TS (Q(1,5p)OOH → Q(1,4s)OOH) 64.3 60.1 60.0 Product conformer, Q(1,4s)OOH 20.1 18.9 17.2 Ethene_HCOOH_OH_reactant 20.7 19.3 18.5 TS (Q(1,5p)OOH → ethene, HCOOH, OH) 49.3 44.8 42.6 Ethene_HCOOH_OH_complex -28.9 -33.1 -41.7 Ethene + HCOOH + OH -1.2 -6.5 -27.7 Cyclopropanol_OH_reactant 20.2 18.8 18.1 TS (Q(1,5p)OOH → cyclopropanol +OH) 66.9 64.2 64.1 Cyclopropanol + OH 36.8 34.5 23.3 Oxetanyl_OH_reactant 18.2 17.0 16.7 TS (Q(1,5p)OOH → oxetane + OH) 44.8 42.2 42.3 Oxetanyl_OH_complex -9.8 -10.5 -12.2 2-Hydroxy oxetane + OH -2.2 -4.3 -13.6

331

B3LYP/6-31+G** ΔH (0 K) ΔH (298 K) ΔG (298 K) Peroxy propanol (tGg’ conformer) 0.0 0.0 0.0 Peroxy propanol, 1,3-H reactant conformer 2.4 2.3 2.3 Peroxy propanol, 1,4-H reactant conformer 1.3 1.3 1.51 Peroxy propanol, 1,5-H reactant conformer 0.3 0.2 0.17 TS_1,3─H-atom transfer 46.4 42.4 42.8 TS_1,4─H-atom transfer 37.2 32.8 33.9 TS_1,4─elimination 30.5 26.5 26.6 TS_1,5_H-atom transfer 30.0 25.5 27.1 Propanoic acid + OH -37.7 -39.0 -42.4 Propanoic acid + OH (Separated products) -32.0 -34.7 -45.0

Propenol + HO2 (product) 24.7 22.1 10.1

Propanal + HO2 (product) 15.6 12.7 0.3

Q(1,4s)OOH 19.1 17.6 16.2 Cis_oxirane_reactant 17.0 15.5 14.2 TS (Q(1,4s)OOH → cis-oxirane + OH 24.9 22.0 20.3 Cis_oxirane_complex -2.7 -4.5 -5.8 Cis-1-hydroxy-2-methyl oxirane 5.0 1.9 -7.6 Trans_oxirane_reactant 16.9 15.4 14.1 TS (Q(1,4s)OOH → trans-oxirane + OH 24.3 21.5 20.6 Trans-oxirane + OH (complex) -3.2 -5.0 -6.3 Trans_oxirane + OH 4.3 1.3 -8.1 Propenol_reactant 16.6 15.1 13.8 TS (Q(1,4s)OOH → propenol + OH) 29.2 26.9 26.3

Propenol + HO2 (complex) 16.9 15.8 12.1

Propenol + HO2 24.7 22.1 10.1

Propanal + HO2 15.6 12.7 0.3

Table 7.6. Enthalpies and free energies of reaction (kcal/mol) calculated at the B3LYP/6-31+G** level of theory at 0 K and 298 K for the unimolecular decomposition of propan-1-ol-1-peroxy radical. See Figure 7.4 for the corresponding pathways. Table 7.6 continued on page 328.

332

Table 7.6, continued.

B3LYP/6-31+G** ΔH (0 K) ΔH (298 K) ΔG (298 K) Q(1,5p)OOH 22.8 21.2 20.8 TS (Q(1,5p)OOH → Q(1,4s)OOH) 62.8 58.5 58.2 Product conformer, Q(1,4s)OOH 20.1 18.6 17.2 Ethene_HCOOH_OH_reactant 21.7 20.2 19.3 TS (Q(1,5p)OOH → ethene, HCOOH, OH) 39.1 35.6 34.3 Ethene_HCOOH_OH_complex -17.7 -20.3 -28.2 Ethene + HCOOH + OH -7.4 -12.5 -33.5 Cyclopropanol_OH_reactant 21.4 19.8 19.2 TS (Q(1,5p)OOH → cyclopropanol +OH) 59.2 56.4 56.3 Cyclopropanol + OH 37.1 34.6 23.6 Oxetanyl_OH_reactant 19.6 18.2 17.9 TS (Q(1,5p)OOH → oxetane + OH) 38.4 35.9 35.9 2-Hydroxy oxetane + OH 3.3 1.1 -7.9

333

45.7 46.4 50.7

1,3-H atom transfer 36.3 30.5 39.3 36.0 37.2 40.1

1,4-H atom transfer/ 1,4-H atom transfer/ HO2 elimination

29.3 30.0 32.8

1,5-H atom transfer

CBS-QB3 0.0 B3LYP 0.0 mPW1K 0.0

Figure 7.5. Initial kinetic barriers for pathways available to propan-1-ol-1-peroxy radical, expressed in terms of ΔH0K (kcal/mol), calculated via CBS-QB3 and DFT (B3LYP/6-31+G** and mPW1K/6-31+G**) methods. Given the complexity of this reactive species, rotations occur prior to these isomerizations, yielding slightly different reactant conformers in each case; however, these result in minimal changes to the enthalpy and are excluded here.

334 Propan-1-ol-1-peroxy radical can revert to reactants or undergo several internal

reactions: 1,3-H transfer; 1,4-H transfer; 1-4-H transfer concurrent with hydroperoxyl

• (HO2 ) elimination; and 1,5-H transfer. Of the four progressive reactions, both 1,3-H

transfer and 1,4-H transfer concurrent with elimination (hereafter referred to as 1,4-

elimination) were shown to yield distinct combustion products: 1,3-H transfer proceeded

with a barrier height of 45.7 kcal/mol to yield propanoic acid and hydroxyl radical, while

the 1,4-elimination pathway incurred a barrier of 36.3 kcal/mol to give propenol and

hydroperoxyl radical. In both cases, displacement of the transition state’s geometry

towards the products yielded a complex of the two products; interestingly, in the case of

propanoic acid and hydroxyl radical, this complex was more stable than the infinitely

separated products, while this trend was reversed for propenol and hydroperoxyl radical.

Both the 1,4-H transfer and 1,5-H transfer resulted in hydroperoxyalkyl radicals; their

barrier heights were 36.0 kcal/mol and 29.3 kcal/mol, respectively. (For all four of the

H-atom transfer pathways, the actual reactant conformer was slightly higher in energy

than the global minimum; these values are also included in the relevant tables, but were

negligible in all cases.)

These findings were in keeping with our earlier work on n-propylperoxy radical.

The 1,5-H transfer was seen to have a notably lower barrier, in both cases, due to the facile six-membered transition state through which it could occur. Both processes involving 1,4-H transfer had comparable barrier heights, with 1,4-elimination occurring at a slightly higher cost. The 1,3-H transfer demonstrated a high barrier, due to the large amount of ring strain invoked by the structural restraints.

335 Also, we can compare the quantitative effect of the additional hydroxyl functionality on the reaction energetics relative to those of n-propylperoxy radical. As

seen in Table 7.7, propan-1-ol-1-peroxy radical has barriers to the H-atom transfers that

were roughly 5−6 kcal/mol higher than those of n-propylperoxy radical, in all cases. This could be attributed to either a higher-energy transition state or a stabilized reactant, in the case of propan-1-ol-1-peroxy radical; it is likely that the hydroxyl group would lead to increased steric effects in the transition state, and could also have the potential for destabilizing inductive effects. A few slight differences were observed. In particular, n-

propylperoxy radical demonstrated a slightly increased preference for 1,5-H transfer over

the other kinetic pathways than its functionalized derivative. For n-propylperoxy radical,

1,5-H transfer was 7.0 kcal/mol more favorable than the next likeliest pathway (1,4-H

• transfer/HO2 elimination), while for propan-1-ol-1-peroxy radical, this preference was

6.7 kcal/mol, and the next likeliest route was 1, 4-H transfer.

336

Propan-1-ol-1-peroxy Radical n-Propylperoxy Radicala

ΔH0K ΔG298K ΔH0K ΔG298K peroxy propanol, lowest energy conformer 0.0 0.0 n-propylperoxy radical 0.0 0.0

TS_1,3─H-atom transfer 45.7 42.0 TS (1,3) 40.9 41.0 propanoic acid + OH (product) −34.1 −47.1 propanal + OH −25.2 −34.2

TS_1,4─H transfer/HO2 elimination 36.3 32.5 TS (1,4_elim) 30.9 31.2 propenol + HO2 (product) 4.3 4.4 propene + HO2 18.2 4.6

TS_1,4─H-atom transfer 36.0 32.9 TS (1,4) 32.1 32.8

Q(1,4s)OOH 16.8 14.2 Q(1,4s)OOH TS(Q(1,4s)OOH → cis-oxirane + OH) 26.6 22.7 TS(Q(1,4s)OOH → oxirane + OH) 25.5 24.7 cis-1-hydroxy-2-methyl-oxirane + OH −3.0 −10.4 methyl oxirane + OH −4.0 −12.1 TS(Q(1,4s)OOH → trans-oxirane + OH) 26.8 22.6 trans-1-hydroxy-2-methyl-oxirane + OH 1.7 -10.7

TS(Q(1,4s)OOH → propenol + HO2 33.2 30.3 TS(Q(1,4s)OOH → propene + HO2 28.7 28.1 propenol + HO2 27.9 13.4 propene + HO2 18.2 7.6

TS_1,5─H-atom transfer 29.3 26.6 TS (1,5) 23.9 24.9

Q(1,5p)OOH 19.8 18.0 Q(1,5p)OOH TS (Q(1,5p)OOH → Q(1,4s)OOH) 52.3 17.1 TS (Q(1,5p)OOH → Q(1,4s)OOH) 53.9 53.7 Q(1,4s)OOH, via Q(1,5p)OOH 15.0 15.5 Q(1,4s)OOH, via Q(1,5p)OOH 13.4 12.3 TS (Q(1,5p)OOH → ethene + HCOOH + TS (Q(1,5p)OOH → ethene + CH2O + OH) 34.6 37.5 OH) 43.2 42.1 ethene + + OH −18.2 −2.0 ethene + formaldehyde + OH 3.6 −14.7 TS (Q(1,5p)OOH → cyclopropanol + TS (Q(1,5p)OOH → cyclopropane + HO2) 57.6 60.5 HO2) 54.2 53.6

cyclopropanol + HO2 33.6 25.6 cyclopropane + HO2 26.9 17.8 TS (Q(1,5p)OOH → hydroxy oxetane + OH) 32.9 35.5 TS (Q(1,5p)OOH → oxetane + OH) 35.6 35.6 hydroxy oxetane + OH −2.2 −10.6 oxetane + OH −0.1 −7.9 aEnthalpies and energies from Ref. 26.

Table 7.7. Comparison of unimolecular decompositions of respective peroxy radicals. CBS-QB3 enthalpies of reaction at 0K and free energies of reaction at 298 K, relative to the global minimum of the respective starting peroxy radical.

337

Both the DFT methods matched the qualitative predictions of the CBS-QB3 calculations,

in supporting 1,5-H atom transfer as the likeliest isomerization route and in the relative

trend of barrier heights: 1,5-H transfer < 1,4-elimination < 1,4-H transfer < 1,3-H

transfer. Quantitatively, compared to the CBS-QB3 results, mPW1K/6-31+G** resulted

in energies that were consistently around 5 kcal/mol too high, while B3LYP/6-31+G**

predicted a relatively low barrier height (30.5 kcal/mol) for 1,4-elimination, but was

otherwise comparable to CBS-QB3 in its calculations.

7.3.3. Subsequent decompositions. Although the QOOH radicals have the capacity to

add a second equivalent of O2 and undergo chain-branching reactions, we focus here on

their unimolecular decompositions. A qualitative view of the decompositions is available

in Figure 7.6, and these pathways have been tabulated (Tables 7.4─ 7.6). In particular,

some reaction steps take place via different reactant conformers and/or result in isolable complexes, rather than immediately proceeding from reactant to transition state to

infinitely-separated products. In these cases, all of the relevant data are incorporated in

Tables 7.4─ 7.6, even while the pictured geometries represent a simplified view.

338

Figure 7.6. Overall potential energy surface for the decomposition of propan-1-ol-1- peroxy radical using ΔH0K (kcal/mol) values calculated at the CBS-QB3 level of theory (qualitative representation − not drawn to scale). Neither reactant conformers nor product complexes were included, to ease interpretation; all of these values are included in Tables 7.3, 7.4 and 7.5. Pathways for the Q(1,5p)OOH isomer are denoted in blue; pathways for the Q(1,4s)OOH isomer, in green; the 1,4-H atom transfer concurrent with • HO2 elimination, in purple; the 1,3-H atom transfer, in red.

Data for the decomposition of the n-propylperoxy radical and 1-peroxy propanol are again presented in tandem (Table 7.7), for the decompositions of the Q(1,4s)OOH and

Q(1,5p)OOH radicals of propan-1-ol-1-peroxy radical. The Q(1,4s)OOH radical of propan-1-ol-1-peroxy radical had the capability to undergo β-scission to form propenol

• and HO2 or to cyclize to form either cis- or trans- 1-hydroxy-2-methyl oxirane, while

339 losing HO• radical. Both cyclization pathways had comparable barrier heights, with the

cis cyclization being slightly lower (26.6 kcal/mol) than the trans route (26.8 kcal/mol);

• these reactions were both slightly exothermic overall, as cis-oxirane and HO2 formed with a reaction enthalpy of –5.7 kcal/mol, and trans-oxirane demonstrated a reaction enthalpy of –5.9 kcal/mol. Scission to propenol and hydroperoxyl occurred with a barrier height of 33.2 kcal/mol and a reaction enthalpy of 27.9 kcal/mol. (If tautomerization could occur to form propanal and hydroperoxyl, the subsequent reaction enthalpy would decrease further to 19.7 kcal/mol; however, in the gas phase, the relevant uncatalyzed process would experience a substantial energy barrier.39)

B3LYP/6-31+G** and mPW1K/6-31+G** methods replicated the qualitative

trends seen by CBS-QB3: cyclization to yield trans-1-hydroxy-2-methyl oxirane and HO• was most favorable, followed by cyclization to cis-1-hydroxy-2-methyl oxirane and HO•,

• then scission to hydroperoxyl (HO2 ) radical and propenol. Quantitatively, the barrier

heights predicted by mPW1K were 6−8 kcal/mol too high, while those predicted by

B3LYP were generally within 2−3 kcal/mol of the CBS-QB3 values.

Compared to n-propylperoxy radical, the general trend of cyclization over

scission was repeated. The preference for cyclization is roughly 3 kcal/mol in n-

propylperoxy’s Q(1,4s)OOH isomer; with the Q(1,4s)OOH isomer of propan-1-ol-1-

peroxy radical, that partiality is more pronounced, around 6 kcal/mol. The kinetic

• barriers to cyclization are analogous for both species; scission to propenol and HO2 , for

• the hydroxy derivative, costs around 5 kcal/mol more than scission to propene and HO2 , in the non-functionalized, parent propylperoxy species. The enthalpies of reaction were

340 comparable: formation of the oxirane derivatives was slightly exothermic, and formation

of the scission products was slightly endothermic, overall.

The Q(1,5p)OOH radical likewise demonstrated several potential pathways. It

could undergo scission to yield ethene, formic acid, and hydroxyl radical; also, a

subsequent H-atom transfer was possible and could yield the Q(1,4s)OOH isomer,

discussed previously. Additionally, two cyclizations were possible: to cyclopropanol, while losing hydroperoxyl radical; or to hydroxy-oxetane, while losing hydroxyl radical.

Except for the scission process, all three steps involved cyclic transition states, so that their enthalpies of activation could be related back to the amount of ring strain present in these transition states. Cyclization to cyclopropanol concurrent with loss of hydroperoxyl occurred with a reaction barrier of 57.6 kcal/mol and an enthalpy of reaction of 33.6 kcal/mol. 1,2-H transfer to yield Q(1,4s)OOH involved a similarly high enthalpic cost

(52.3 kcal/mol) but a lower enthalpy of reaction (15.0 kcal/mol).40 The four-membered

cyclic transition state afforded by 2-hydroxy-oxetane formation and HO• loss saw a substantially lower kinetic barrier (32.9 kcal/mol). As seen in the initial H-atom transfers, a direct correlation again could be made between the size of the cyclic transition state and the height of the relevant kinetic barrier. Finally, scission to ethene, formic acid, and hydroxyl radical occurred with a barrier height of 34.6 kcal/mol; the dissociative nature of this reaction led to an overall exothermicity (−2.2 kcal/mol), which was unsurprisingly more pronounced in the free energy data.

The DFT methods repeated most of the trends for Q(1,5p)OOH decomposition.

Cyclization to 2-hydroxy oxetane, concurrent with loss of HO• radical, demonstrated the

341 lowest kinetic barrier; scission to ethene, formic acid, and HO• radical was the second

most energetically-favorable process. B3LYP/6-31+G** predicted that cyclopropanol formation with loss of HO• radical would experience a lower barrier than 1,2-H transfer,

which reversed the trend seen by CBS-QB3 and mPW1K.

Comparing the chemistry of propan-1-ol-1-peroxy radical to that of n-

propylperoxy radical, the same preferential trends are seen: cyclization to an oxetanyl species with loss of HO• is favored, followed by scission to ethene and various

oxygenated species, then 1,2-H transfer to form the Q(1,4s)OOH isomer, then cyclization

• to a three-membered ring with loss of HO2 . Generally, n-propylperoxy radical saw

higher kinetic barriers to its reaction than propan-1-ol-1-peroxy radical, except in the case

•, of cyclization to cyclopropanol and HO2 for propan-1-ol-1-peroxy radical. Looking at

the reaction thermodynamics, the oxetanyl and scission pathways were more exothermic

in propan-1-ol-1-peroxy radical, while the cyclopropanol and isomerization pathways

were more endothermic, than in the n-propylperoxy system. Most enthalpies of reaction

were comparable, except in the case of scission to ethene, formic acid, and HO•, which

was quite exothermic (-18.2 kcal/mol) for propan-1-ol-1-peroxy radical, while the

corresponding scission to ethene, formaldehyde, and HO• was slightly endothermic (3.6

kcal/mol) for n-propylperoxy radical.

7.3.4. Product distribution and atmospheric implications

The chief difference in qualitative product distribution, between propan-1-ol-1-

peroxy radical and n-propylperoxy radical, involved increased oxygen content in the

342 products of the former, relative to the latter. Propanal was formed rather than propene,

via the 1,4-elimination process. Carbonyl compounds react slightly less readily with the

• • prevalent atmospheric species HO and NO3 radicals than do the corresponding

hydrocarbons.41,42 For instance, the rate coefficient for propene and HO• is 26.3 x 1012 cm3/(molecule-s); while the rate coefficient for propanal and HO• is 20 x 1012

3 • • cm /(molecule-s). With NO3 , a similar effect is seen: propene and NO3 demonstrate a

–15 3 • rate coefficient of 9.49 x 10 cm /(molecule-s), while propanal and NO3 show a rate

coefficient of 6.5 x 10-15 cm3/(molecule-s). Additionally, the difference in product

distribution could have effects on the amount and identity of the relevant products.

Aldehydes are known to be precursors to peroxyacylnitrates (PANs), which demonstrate

lachrymatory and mutagenic effects. PANs are long-lived so can transport other radicals

away from their points of origin, thus increasing the lifetime and impact of these species;

biofuels could thus have a negative impact with respect to some atmospherically- detrimental species, even while decreasing the concentrations of others.

• 1,4-H transfer without concurrent HO2 elimination yielded a variety of other oxygenated products, including propenol and two isomers of 1-hydroxy-2-methyl oxirane. Enols have recently been implicated as intermediates in high-temperature hydrocarbon combustion;43 it is hypothesized that they will primarily react via H-atom

abstraction and alkyl-radical addition to the double bond, yielding larger hydrocarbon

products (perhaps with implications for soot production). These processes would be

involved in propenol’s chemistry. Also, although tautomerization is unlikely at

tropospheric chemistry temperatures, propenol could lead to aldehydes via other

343 pathways. With respect to the 1-hydroxy-2-methyl oxirane isomers, it seems unlikely

that the would figure directly into their chemistry, given its strong O−H

bonds; thus, the atmospheric pathways available to ethers (primarily H-atom abstraction

via HO• radical) are likely to remain factors in this case. Since ethers generally see

increased reactivity at the positions ortho to the oxygen atom, it is likely that the hydroxy

carbon, situated as it is between two oxygen atoms, will see more pronounced reactivity.

For propan-1-ol-1-peroxy radical, the 1,3-H transfer yielded propanoic acid and

HO•. Carboxylic acids have implications in acid rain formation and can also serve as

• • sources of HO and HO2 ; thus, they are of environmental interest. However, since the

barrier height to this process is so prohibitive, it is unlikely to contribute to the overall

chemistry of propan-1-ol-1-peroxy radical. Indeed, atmospheric carboxylic acids are

generally formed as ozonolysis products, via rearrangements of carbonyl oxides.44

1,5-H transfer in propan-1-ol-1-peroxy radical could lead to a variety of products relative to the same process in n-propylperoxy radical: cyclopropanol instead of

cyclopropane; formic acid instead of formaldehyde; functionalized oxetane rather than

oxetane alone. Cyclopropanol is not expected to demonstrate significantly different reactivity than its counterpart via the n-propylperoxy processes; the barrier height to this step implies that it is unlikely to even occur. At most, the presence of the hydroxy group could lead to preferential radical formation in terms of H-atom abstraction, which in turn

could lead to characteristic decomposition products. Hydroxy-oxetane formation and scission to ethene, formic acid, and HO• demonstrated the lowest barriers available to the

Q(1,5p)OOH isomer of propan-1-ol-1-peroxy radical; these processes correspond to

344 formation of oxetane and scission to ethene, formaldehyde, and HO• for the Q(1,5p)OOH isomer of n-propylperoxy radical. The most pronounced difference is the formation of formic acid rather than formaldehyde; as noted above, carboxylic acids do not tend to form directly in the atmosphere, so this is a notable pathway. It is important to remember, though, that although the Q(1,5p)OOH isomer readily forms from its peroxy precursor, the barriers to all subsequent processes are higher than those for the corresponding steps available to the Q(1,4s)OOH isomer, and it is likely that none of the pathways available to the Q(1,5p)OOH isomer will have a significant effect on product distribution.

Given the overall similarity between the chemistry of n-propylperoxy radical and propan-1-ol-1-peroxy radical, it seems likely that propan-1-ol-1-peroxy radical will decompose by 1,4-H transfer concurrent with hydroperoxyl radical elimination, which is the pathway favored by its hydrocarbon counterpart. For propan-1-ol-1-peroxy radical,

• this would result in the formation of propenol and HO2 . Additionally, the 1,4-H transfer pathway was more competitive in propan-1-ol-1-peroxy radical than in n-propylperoxy radical. In n-propylperoxy radical, 1,4-H transfer saw a substantially higher barrier than

1,4-H transfer concurrent with elimination of hydroperoxyl radical; in propan-1-ol-1- peroxy radical, the barrier to the concerted pathway was actually 0.6 kcal/mol less favorable than the H-atom transfer alone. Thus, the cyclic ethers accessible via 1,4-H transfer could contribute to the product distribution.

345 7.4. Conclusions

In the course of this work, we have built on our previous study of the n-

propylperoxy radical to explore the effect of the hydroxyl functionality on the reaction pathways commonly seen in low-temperature oxidation. In 1-propanol, the most likely

peroxy radical, derived from initial homolytic C−H bond cleavage and subsequent O2 addition, was identified, and the multiple conformers of propan-1-ol-1-peroxy radical were examined, in order to determine the global minimum. It was seen that several conformers existed in appreciable amounts at room temperature; their energies are within

2 kcal/mol, so it is expected that rotamer interconversion will not compete with the decomposition pathways of this species.

The initial H-atom transfers available to propan-1-ol-1-peroxy radical were modeled via CBS-QB3 and DFT calculations. As was the case with the alkylperoxy analogue, 1,5-H transfer occurred via a facile, six-membered transition state and demonstrated a low barrier height; barrier heights for the H-atom transfers increased proportionally to the amount of strain in the transition state. One difference between n-

propylperoxy radical and propan-1-ol-1-peroxy radical was seen in the relative preference

• for 1,4-H transfer concurrent with HO2 elimination and 1,4-H transfer alone. With n-

propylperoxy radical, the former process was favored; however, with propan-1-ol-1-

peroxy radical, the latter process was favored. In each case, the two processes were

quantitatively similar (with propan-1-ol-1-peroxy radical, the preference for 1,4-H

transfer was only 0.3 kcal/mol). Moreover, this was one of the few steps in which the

CBS-QB3 predictions varied with respect to the DFT work: both B3LYP and mPW1K

346 predicted that concerted 1,4-elimination would be significantly more favorable than 1,4-

H transfer alone.

The subsequent decomposition pathways of each of the resulting isomers were

modeled. The basic trends of the reactivity of n-propylperoxy radical were replicated in

propan-1-ol-1-peroxy radical. The pathways available to the Q(1,4s)OOH isomer of propan-1-ol-1-peroxy radical demonstrated the lowest barriers of the subsequent routes and are thus postulated to contribute to the actual decomposition of the peroxy radical species. Qualitatively, the main difference in product distribution involved increased oxygen content in the decomposition products: propenol; cis- and trans-1-hydroxy-2-

• • methyl-oxirane, HO , and HO2 were commonly seen. In particular, propenol and the 1-

hydroxy-2-methyl-oxiranes can lead to relatively unusual chemistry in the troposphere.

Finally, while the species examined in this study were small, and thus CBS-QB3

calculations were not prohibitive, it was notable that B3LYP and mPW1K both replicated

nearly all of the trends predicted by the comprehensive method, at substantially lower

computational cost. In particular, B3LYP/6-31+G** produced results that were quantitatively comparable to those of the CBS-QB3 method. For the larger species that constitute true biomass compounds, hybrid DFT methods may prove invaluable in providing economical results. Further work in this area could extend the size of either the of interest (perhaps exploring the effect of the ester functionality) or the length of the alkyl chain, as both areas would have implications for biofuel chemistry.

Thus, while the structural addition of the hydroxyl group to n-propylperoxy

radical is a relatively minor change, it did affect both the quantitative and qualitative

347 decompositions of the relevant species. Current experimental surveys of biomass compounds propose key differences in the combustion of these species relative to hydrocarbon chemistry, both beneficial and detrimental, and these results support those findings. For instance, while increased oxygen content in these fuels will lead to lesser carbon monoxide formation, the various products formed could lead to increased soot production and/or acidic content. The generation of these fuels presents substantial challenges for mechanistic and combustion chemistry, as do the decompositions of the various reactive intermediates afforded to these processes.

348 References for Chapter 7

1 (a) Curran, H.J.; Gaffuri, P.; Pitz, W.J.; Westbrook, C.K. Combust. Flame 1998, 114,

149-177. (b) Curran, H.J.; Gaffuri, P.; Pitz, W.J.; Westbrook, C.K. Combust. Flame

2002, 129, 253-280. (c) Curran, H. J.; Pitz, W. J.; Westbrook, C. K.; Callahan, C. V.;

Dryer, F. L. Proc. Combust. Inst. 1998, 27, 379-387; (d) Huang, Y.; Sung, C. J.; Eng, J.

A. Combust. Flame 2004, 139, 239-251; (e) Huang, Y.; Sung, C. J.; Eng, J. A. Proc.

Combust. Inst. 2002, 29, 759-766.

2 Smith, G. P.; Golden, D. M.; Frenklach, M.; Moriarty, N. W.; Eiteneer, B.; Goldenberg,

M.; Bowman, C. T.; Hanson, R. K.; Song, S.; Gardiner, W. C., Jr.; Lissianski, V. V.; Qin,

Z. http://www.me.berkeley.edu/gri_mech/.

3 (a) Freedman, B.; Pryde, E. H. Proc. Int’l. Conf. on Plant and Vegetable Oils as Fuels,

1982, 117–22. (b) Lipinsky, E. S.; McClure, T. A.; Wagner, C. K., Kresovich, S., Otis, J.

L.; Applebaum, H. R. Proceedings of the International Conference on Plant and

Vegetable Oils as Fuels, 1982. 1–10. (c) Fuls, J.; Hugo, F. J. C. Third International

Conference On Energy Use Management, 1981. 1595–602. (d) Demirba, A. Energ.

Convers. Manage. 2002, 43, 2349-2356.

4 Marinov, N. M. Int. J. Chem. Kinet. 1999, 31, 183-220.

5 Cavalli, F.; Geiger, H.; Barnes, I.; Becker, K.H. Environ. Sci. Technol. 2002, 36, 1263-

1270.

6 McEnally, C.S.; Pfefferle, L.D. Proc. Combust. Inst. 30:1363-1370 (2005).

7 Fisher, E.M.; Pitz, W.J.; Curran, H.J.; Westbrook, C.K. Proc. Combust. Inst. 2000, 28,

1579-1586.

349

8 Gail, S.; Thomson, M. J.; Sarathy, S. M.; Syed, S.A.; Dagaut, P.; Dievart, P.; Marchese,

A. J.; Dryer, F. L. Proc. Combust. Inst. 2007, 31, in press.

9 Good, D. A.; Francisco, J. S. J. Phys. Chem. A 2000, 104, 1171-1185.

10 (a) Metcalfe, W. K.; Dooley, S.; Curran, H. J.; Simmie, J. M.; El-Nahas, A. M.;

Navarro, M. V.2007, 111, 4001-4014; (b) El-Nahas, A. M.; Navarro, M. V.; Simmie, J.

M.; Bozzelli, J. W.; Curran, H. J.; Dooley, S.; Metcalfe, W. J. Phys. Chem. A. 2007, 111,

3727-3739.

11 (a) Schwab, A. W.; Bagby, M. O.; Freedman, B. Fuel 1987, 66, 1372-1378; (b)

Schwab, A. W.; Dykstra, G. J.; Selke, E.; Sorenson, S. C.; Pryde, E. H. JOACS 1988, 65,

1781-1786.

12 Graboski, M. S.; McCormick, R. L. Prog. Energy Combust. Sci. 1998, 24, 125-164.

13 Srivastava, A.; Prasad, R. Renew. Sust. Energ. Rev. 2000, 4, 111-133.

14 Lin, C.-Y.; Lin, H.-A.; Hung, L.-B. Fuel 2006, 85, 1743-1749.

15 Ma, F.; Hanna, M. A. Bioresour. Technol. 1999, 70, 1-15.

16 Bozzelli, J. W.; Sheng, C. J. Phys. Chem. A 2002, 106, 1113-1121.

17 Walker, R. W.; Morley, C. Basic Chemistry of Combustion. In Low-Temperature

Combustion and Autoignition; Willing, M. J., Ed.; Elsevier: Amsterdam; New York,

1997; pp 1-124.

18 (a) Bozzelli, J. W.; Pitz, W. J. Twenty-Fifth Symposium (International) on Combustion;

The Combustion Institute: Pittsburgh, PA, 1994; 783-791. (b) Bozzelli, J. W.; Sheng, C.

J. Phys. Chem. A 2002, 106, 1113-1121.

350

19 Finlayson-Pitts, B. J.; Pitts, J. N. Chemistry of the Upper and Lower Atmosphere:

Theory, Experiments, and Applications; Academic Press, San Diego, CA, 2000.

20 (a) Wagner, A. F.; Slagle, I. R.; Sarzynski, D.; Gutman, D. J. Phys. Chem. 1990, 94,

1853. (b) Kaiser, E. W.; Wallington, T. J.; Andino, J. M. Chem. Phys. Lett. 1990, 168,

309. (c) Kaiser, E. W.; Rimai, L.; Wallington, T. J. J. Phys. Chem. 1989, 93, 4094. (d)

Kaiser, E. W.; Lorkovic, I. M.; Wallington, T. J. J. Phys. Chem. 1990, 94, 3352. (e)

Dobis, O.; Benson, S. W. J. Am. Chem. Soc. 1993, 115, 8798. (f) Kaiser, E. W. J. Phys.

Chem. A 2002, 106, 1256. (g) Kaiser, E. W. J. Phys. Chem. 1995, 99, 707. (h) Clifford,

E. P.; Farrell, J. T.; DeSain, J. D.; Taatjes, C. A. J. Phys. Chem. A 2000, 104, 11549. (i)

Bozzelli, J. W.; Dean, A. M. J. Phys. Chem. 1990, 94, 3313. (j) Quelch, G. E.; Gallo, M.

M.; Schaefer, H. F., III. J. Am. Chem. Soc. 1992, 114, 8239. (k) Ignatyev, I. S.; Xie, Y.;

Allen, W. D.; Schaefer, H. F., III. J. Chem. Phys. 1997, 107, 141. Stark, M. S. J. Am.

Chem. Soc. 2000, 122, 4162. (l) Chen, C.-J.; Bozzelli, J. W. J. Phys. Chem. A 2000, 104,

4997. (m) Rienstra-Kiracofe, J. C.; Allen, W. D.; Schaefer, H. F., III. J. Phys. Chem. A

2000, 104, 9823. (n) Miller, J. A.; Klippenstein, S. J.; Robertson, S. H. Proc. Combust.

Inst. 2000, 28, 1479. (o) Bozzelli, J. W.; Sheng, C. J. Phys. Chem. A 2002, 106, 1113.

21 Ignatyev, I. S.; Xie, Y.; Allen, W. D.; Schaefer, H. F., III. J. Chem. Phys. 1997, 107,

141-155.

22 (a) Naik, C.; Carstensen, H.-H.; Dean, A. M. Proceedings of the Third Joint Meeting of

the U.S. Sections of the Combustion Institute, Chicago, IL, 2003. (b) Chan, C.-J.;

Hamilton, I. P.; Pritchard, H. O. Faraday Trans. 1998, 94, 2303-2306.

351

23 DeSain, J. D.; Taatjes, C. A.; Miller, J. A.; Klippenstein, S. J.; Hahn, D. K. Faraday

Discuss. 2001, 119, 101-120.

24 (a) DeSain, J. D.; Klippenstein, S. J.; Miller, J. A.; Taatjes, C. A. J. Phys. Chem. A

2003, 107, 4415-4427. (b) DeSain, J. D.; Klippenstein, S. J.; Miller, J. A.; Taatjes, C. A.

J. Phys. Chem. A 2004, 108, 7127-7128.

25 DeSain, J. D.; Clifford, E. P.; Taatjes, C. A. J. Phys. Chem. A 2001, 105, 3205-3213.

26 Merle, J. K.; Hayes, C. J.; Zalyubovsky, S. J.; Glover, B. G.; Miller, T. A.; Hadad, C.

M. J. Phys. Chem. A. 2005, 109, 3637-3646.

27 (a) Wagner, A. F.; Slagle, I. R.; Sarzynski, D.; Gutman, D. J. Phys. Chem. 1990, 94,

1853; (b) Kaiser, E. W.; Wallington, T. J.; Andino, J. M. Chem. Phys. Lett. 1990, 168,

309; (c) Kaiser, E. W.; Rimai, L.; Wallington, T. J. J. Phys. Chem. 1989, 93, 4094; (d)

Kaiser, E. W.; Lorkovic, I. M.; Wallington, T. J. J. Phys. Chem. 1990, 94, 3352; (e)

Dobis, O.; Benson, S. W. J. Am. Chem. Soc. 1993, 115, 8798; (f) Kaiser, E. W. J. Phys.

Chem. A 2002, 106, 1256; (g) Kaiser, E. W. J. Phys. Chem. 1995, 99, 707; (h) Clifford,

E. P.; Farrell, J. T.; DeSain, J. D.; Taatjes, C. A. J. Phys. Chem. A 2000, 104, 11549.

28 Zalyubovsky, S. J.; Glover, B. G.; Miller, T. A.; Hayes, C. J.; Merle, J. K.; Hadad, C.

M. J. Phys. Chem. A 2005, 109, 1308-1315.

29 Vereecken, L.; Peters, J. J. Chem. Phys. 2003, 119, 5159-5170.

30 (a) Parr, R. G.; Yang, W. Density Functional Theory in Atoms and Molecules; Oxford

University Press: New York, 1989. (b) Labanowski, J. W.; Andzelm, J. Density

Functional Methods in Chemistry; Springer: New York, 1991.

352

31 (a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648; (b) Becke, A. D. Phys. Rev. A. 1988,

38, 3098.

32 Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B 1998, 37, 785-789.

33 Lynch, B. J.; Fast, P. L.; Harris, M.; Truhlar, D. G. J. Phys. Chem. A 2000, 104, 4811.

34 (a) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A., Jr. J. Chem. Phys. 1996,

104, 2598-2619; (b) Petersson, G. A.; Frisch, M. J. J. Phys. Chem. A 2000, 104, 2183-

2190; Montgomery, J. A., Jr.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. A. J. Chem.

Phys. 1999, 110, 2822-2827.

35 Gaussian 03, Revision C.02, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G.

E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.;

Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi,

M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.;

Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.;

Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;

Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;

Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.;

Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain,

M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.;

Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.;

Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham,

M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.;

353

Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A. Gaussian, Inc., Wallingford CT,

2004.

36 Scott, A. P.; Radom, L. J. Phys. Chem. 1996, 100, 16502-16513.

37 (a) Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1989, 90, 2154. (b) Gonzalez, C.;

Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523.

38 Methyl rotation and C−C−O−H rotation both occur with relatively minor energetic

effects, so were not systematically explored as the stated dihedrals were.

39 Taatjes, C.; Hansen, N.; McIlroy, A.; Miller, J.; Senosiain, J.; Klippenstein, S.; Qi, F.;

Sheng, L.; Zhang, Y.; Cool, T.; Wang, J.; Westmoreland, P.; Law, M.; Kasper, T.;

Kohse-Hoinghaus, K. Science 2005, 308, 1887-1889.

40 We note that the conformer of Q(1,4s)OOH formed via this step different than that

formed via direct 1,4-H atom transfer in the initial 1-peroxy propanol; hence, the energy

of the two conformers differs by ~1.5 kcal/mol.

41 Atkinson, R.; Arey, J. Chem. Rev. 2003, 103, 4605-4638.

42 (a) Atkinson, R. J. Phys. Chem. Ref. Data 1997, 26, 215; (b) Calvert, J. G.; Atkinson,

R.; Kerr, J. A.; Madronich, S.; Moortgat, G. K.; Wallington, T. J.; Yarwood, G. The

Mechanisms of Atmospheric Oxidation of the Alkenes; Oxford University Press: New

York, 2000; (c) IUPAC; http://www.iupac-kinetic.ch.cam.ac.uk

43 Taatjes, C. A.; Hansen, N.; McIlroy, A. Miller, J. A.; Senosiain, J. P.; Klippenstein, S.

J.; Qi, F.; Sheng, L.; Zhang, Y.; Cool, T. A.; Wang, J.; Westmoreland, P. R.; Law, M. E.;

Kasper, T.; Kohse-Höinghaus, K. Science, 2005, 308, 1887-1889. (b) Taatjes, C. A.;

354

Hansen, N.; Miller, J. A.; Cool, T. A.; Wang, J.; Westmoreland, P. R.; Law, M. E.;

Kasper, T.; Kohse-Hoinghaus, K. Proc. Combust. Inst. 2005, 30, 43-49.

44 Aplincourt, P.; Ruiz-Lopez, M. F. J. Am. Chem. Soc. 2000, 122, 8990-8997.

355

CHAPTER 8

NON-AQUEOUS SOLVATION OF n-OCTANOL AND ETHANOL: SPECTROSCOPIC AND COMPUTATIONAL STUDIES

Reproduced from J. Phys. Chem. B, 2006, 110, 6325-6331. Copyright 2006, American

Chemical Society.

8.1. Introduction

Intramolecular O−H bond interactions in the condensed phase are highly sensitive to their environment.1,2 It is also well known that intramolecular and intermolecular hydrogen bonding plays an important role in many areas of biology and chemistry.

Arising from the interaction of hydrogen with an electronegative atom, this relatively weak interaction dictates a wide variety of phenomena. In particular, several studies3-10 have focused on the biological relevance of hydrogen bonds, from the conformations of carbohydrates7 to nucleoside acidity,8 and from protein folding to enzyme-substrate binding.3 In addition, hydrogen bonding in atmospheric aerosols plays a role in reactions and structural phenomena of aerosol surfaces.9,10

Alcohols are unusual in that their OH functionality is responsive to neighboring molecules and are therefore a sensitive probe of the solvation environment. The O−H bond in alcohols has been studied using a variety of methods, such as infrared

356 spectroscopy2,11,12 and Raman spectroscopy.2,12 IR and Raman spectroscopic methods are

useful because of their sensitivity in detecting the frequency shifts of the O−H stretch.1,2

Methanol,13 ethanol,14 and butanol14 clusters have been studied extensively,

providing information on cluster structures and hydrogen-bonding strengths. These

studies also provide insight into solute-solvent interactions via an examination of the

OH−stretching vibrational frequency region. Yet, vibrational spectra in the condensed

phase of the hydrogen-bonding region can be complex and difficult to interpret.1,2 More recently however, coherent infrared condensed phase studies11 have revealed

phenol/benzene complexation when benzene was used as the solvent. Two-dimensional

infrared vibrational echo spectra were also used to extract the binding kinetics of the

phenol/benzene complexes.11

Generally, the assignments for the OH frequency regions for condensed phase

alcohols are well established.2 The hydrogen-bonding region of alcohols (~3100-3550

cm-1) typically contains four bands associated with the O−H stretch.2 One band at ~3190

cm-1 was assigned to the alcohol O−H that is a single proton-donor and a double proton-

acceptor.2 Another band centered between ~3300 and ~3400 cm-1 was attributed to the

O−H bond that is a single proton-donor, single proton-acceptor (typical of alcohols that

are part of a linear or cyclic oligomer).2 A third band at ~3500 cm-1 was assigned to the

O−H stretch of a single-proton donating alcohol.2 Bands occurring at ~3600 cm-1 were

attributed to free O−H stretches of alcohols.2

Alcohols in aqueous solution exist in several monomer and aggregate forms. Each

form or complex can be distinguished by deconvolution of the O−H stretching bands,

although this can be experimentally challenging. Studies have been completed on 1-

357 octanol and CCl4 mixtures to elucidate linear dimers and complex structures and to

2 examine the interactions between the CCl4 molecules and 1-octanol. Analysis of the

hydrogen-bonded spectral region revealed that, in liquid octanol, both the hydrophobic

and hydrophilic properties of the molecule determine the octanol aggregate structures, i.e.

the presence of both monomers and moderately-sized aggregates.2 Alcohol aggregates

and monomers also possess O−H bonds, which can be spectroscopically assigned as free

O−H stretches. There are three different types of free O−H bonds: 1) an O−H that is a

double proton acceptor, 2) an O−H that is a single proton acceptor, and 3) an O−H that is not involved in hydrogen bonding.

The O−H spectral bands attributed to linear and cyclic oligomers in the hydrogen-

bonded region (3000-3500 cm-1) as stated above are well separated from the free O−H

stretch of alcohol monomers and aggregates. Examination of the isolated free O−H band

at low alcohol concentrations provides an understanding of the basic intermolecular

interactions between the alcohol and the solvent as shown in a previous study.2

In addition to the experimental work on alcohols, several relevant computational

studies have been undertaken. With regard to the solvents of interest, clusters of benzene

and water have been modeled,15-17 as have benzene-methanol clusters;18-21 both systems

have demonstrated red shifts due to H-bonding between the O−H unit and the π-system

of the benzene ring. Molecular dynamics studies have been completed on the

interactions of methanol and carbon tetrachloride,22 as well as alcohol conformations in

general,23 and various alcohol-halide complexes have been modeled via ab initio and

DFT methods.24 However, less work has been done on monomeric interactions of

alcohols and these solvents. Several approaches to computationally model solvent effects

358 have also been documented.25-26 It has been well established that solution-phase

chemistry differs substantially from the gas-phase chemistry that typically provides the

default setting for most theoretical work.27-31 Studies have shown the ability of various methods, including HF, MP2, and DFT, to accurately model solvated systems and the

effect of solvation on vibrational frequencies, 21,32-36 including that of the O−H stretch.20

In this study, the interactions of n-octanol and ethanol with the organic solvents

benzene, carbon tetrachloride (CCl4), and cyclohexane were investigated with Raman

spectroscopy. Computational studies were completed to provide additional information

on the conformations of the solvent-alcohol complexes. Calculated Raman spectra were

generated for comparison to experiment, and binding energies were determined. The

large number of conformers possible for n-octanol made calculations lengthy; thus,

calculations were completed on ethanol, in the anticipation of using this smaller molecule

as a model system for n-octanol. As will be shown below, the experimental ethanol

spectra are very similar to those of n-octanol.

359

8.2 Experimental and Computational Methods

The experimental work reported in this chapter was completed by Lori M. Levering

and Karen C. Callahan, under the guidance of Dr. Heather C. Allen.

8.2.1. Raman Spectroscopy. Unpolarized Raman spectra were acquired using ~67 mW

for the n-octanol studies and ~47 mW for the ethanol studies from a 532 nm continuous

wave (CW) YAG laser (Spectra-Physics, Millennia II). The backscattered light was

collected with a fiber optic probe (InPhotonics) coupled to the entrance slit of a 500 mm

monochromator (Acton Research, SpectraPro 500i) using a 1200 groove/mm grating. The

slit width was set at 100 μm and the bandpass varied between 3.3 cm-1 (at 3570 cm-1) and

3.2 cm-1 (at 3670 cm-1). The spectra were collected in 90 s exposures to a liquid nitrogen

cooled CCD camera (Roper Scientific, LN400EB, 1340 x 400 pixel array, back-

illuminated and deep depletion CCD). SpectraSense software (Acton Research, version

4.1.9) was used for data collection and display. CCD calibration was completed using the 435.83 nm line of a fluorescence lamp. Calibration of the wavenumber position was done by obtaining a spectrum of crystalline and comparing peak positions

with literature values.37 Built-in algorithms in the software package IGOR (version

4.0.5.1) were used to fit the Raman spectra to Gaussian line shapes.

8.2.2. Chemicals. The solvents, cyclohexane (Acros) and benzene (Aldrich) had a purity

of 99.9%, while CCl4 (Aldrich) had a purity of >99.5%. The alcohols, n-octanol (Fisher)

and ethanol (Aldrich), had a purity of 99.9% and 99.5%, respectively. Anhydrous ethanol

(Fisher) and anhydrous magnesium (Fisher) were also utilized.

360

8.2.3. Computational Methods. All geometry optimizations and vibrational frequency calculations were performed using Gaussian0338 at the Ohio Supercomputer Center. HF

and MP2 computational methods, in conjunction with the 6-31G* basis set,39 were used

to calculate the optimized geometries; single-point energies using these optimized

geometries were then calculated at the B3LYP/6-31+G** level,40-43 using the scf=tight

option. In addition, single-point energy calculations with the more extensive B3LYP/6-

311+G(3df,2p) basis set were completed in order to obtain more quantitatively accurate

energies.44

Vibrational frequencies were calculated for each stationary point to verify their

characterization as minima. The HF and MP2 calculations provided the scaled zero-point

vibrational energies (using factors of 0.9135 and 0.9646, respectively)45a and scaled

frequencies (using factors of 0.89 and 0.9427, respectively),45b as well as thermal and

entropic corrections to the enthalpy and free energy. The thermal and entropic corrections

were obtained from the vibrational frequency calculations, using the unscaled

frequencies. The single-point B3LYP/6-31+G** electronic energies were then used in

tandem with the thermal, entropic, and zero-point vibrational energy (ZPE) corrections

from the HF/6-31G* and MP2/6-31G* calculations to obtain better relative energies for

the various species. The OH-stretching regions of the complexes’ Raman spectra were

calculated computationally in order to compare to the experimental data. The various complexes’ vibrational frequencies were Boltzmann-weighted (based on the B3LYP/6-

31+G** H0 values) and their respective Raman intensities adapted to generate theoretical

spectra for comparison to those obtained experimentally. Solvent effects were included in

361 additional runs by performing polarizable continuum model (PCM)25-26,46-49 energy

calculations at the B3LYP/6-31+G** level of theory, using the respective gas-phase

geometries.

8.3 Results and Discussion

The experimental work reported in this chapter was completed by Lori M. Levering

and Karen C. Callahan, under the guidance of Dr. Heather C. Allen.

8.3.1 Peak assignments. To study the effects of solvent on the free O−H region of n- octanol and ethanol, it was necessary to examine the spectroscopic features of neat n- octanol and ethanol. The Raman spectra of n-octanol and ethanol are shown in Figure

8.1.

Neat Octanol 40000 Neat Ethanol

30000

20000 1 0000 3550 3600 3650 3700

Raman Intensity (a.u.) Raman Intensity 10000

0 Figure 1: Raman spectra of neat octanol and ethanol. 1000 1500 2000 2500 3000 3500 4000 Raman Shift (cm-1)

Figure 8.1. Raman spectra of neat octanol and ethanol. Inset: the free OH region of the spectrum is enlarged.

362 The vibrational assignments are as follows. The region between ~1000 and ~1260 cm-1 contains the bands associated with the C−C and C−O stretches.50,51 The C−O−H bending

-1 51 -1 50 band occurs between ~1200 and ~1450 cm . The CH3 bend occurs at ~1460 cm , and the bands associated with the C−H stretches appear between ~2840 cm-1 and ~3000 cm-

1.51 The region from ~3000 to ~3500 cm-1 contains the bands associated with hydrogen-

bonded O−H stretches,2 and the region between ~3550 and ~3670 cm-1 contains a broad

peak centered at ~3639 cm-1 which is attributed to the alcohol free O−H stretch.2,52,53 In

this work, changes in the free O−H band between ~3550 and ~3670 cm-1 due to the

addition of the inert solvents, such as carbon tetrachloride, cyclohexane, and benzene, are

observed.

The Raman spectra of the free O−H region of neat n-octanol and 0.05x n-octanol in CCl4, cyclohexane, and benzene are shown in Figure 8.2, along with their calculated

fits and component peaks (x = mole fraction). The free O−H band was easily fit to two peaks for each spectrum, consistent with the likelihood of two or more types of free O−H stretches being present. The free O−H peaks of neat n-octanol are positioned at 3636 and

-1 3647 cm . The free O−H peaks of n-octanol in CCl4 and cyclohexane occur at 3635 and

3642 cm-1, and 3642 and 3648 cm-1, respectively. When benzene is used as the solvent,

the free O−H peaks of n-octanol are observed at 3611 and 3632 cm-1. When benzene is

used as the solvent, the free O−H peaks of ethanol occur at 3603 and 3613 cm-1.

The Raman spectra of neat ethanol and 0.05x ethanol in CCl4, cyclohexane, and

benzene are shown in Figure 8.3, along with their calculated fits and component peaks.

The predominant free O−H peak of neat ethanol is positioned at 3639 cm-1. The free O−H

-1 peaks of ethanol in CCl4 occur at 3633 and 3638 cm , and in cyclohexane, at 3640 and

363 3646 cm-1. Comparing the Raman spectra of n-octanol and ethanol reveals a distinct trend: the solvents CCl4 and cyclohexane have a small effect on the free O−H band centered at ~3639 cm-1; however, benzene as a solvent significantly red-shifts this band

by ~30 cm-1. In addition, the free O−H band of both alcohols in benzene is significantly more asymmetric relative to the other solvent/alcohol spectra, suggesting a bimodal distribution of distinctly different complexes or multiple O−H stretching frequencies. The

similarities in the two alcohols, and their spectra, indicate that ethanol can be used as a

model system for n-octanol.

364

Figure 8.2. Experimental Raman spectra of (a) neat n-octanol, and 0.05x n-octanol in (b) benzene, (c) cyclohexane, and (d) carbon tetrachloride, in the 3550 to 3700 cm-1 region of the free O−H stretch. The component peaks are shown in gray and the calculated spectral fit from the component peaks is shown as a line going through the majority of the data points.

365 Figure 8.3. Experimental Raman spectra of (a) neat ethanol, and 0.05x ethanol in (b) benzene, (c) cyclohexane, and (d) carbon tetrachloride, in the 3550-3700 cm-1 region of the free O−H stretch. The component peaks are shown in gray and the calculated spectral fit from the component peaks is shown as a line through the majority of the data points.

8.3.2. Nature of O−H bonds. There are several scenarios that may contribute to the

asymmetric character of the free O−H bands at ~3640 cm-1. There are three types of free

O−H groups that can exist in alcohols: 1) an O−H that is a double proton acceptor, 2) an

O−H that is a single proton acceptor, and 3) an O−H that is not involved in hydrogen bonding. Each of these different types of O−H groups will have a slightly different O−H stretching frequency. The double proton acceptor O−H will have the lowest stretching frequency and the O−H that is not involved in any hydrogen bonding will have the highest stretching frequency. With adequate separation between the different free O−H band frequencies, the free O−H band would be expected to have trimodal character.

However, the Raman spectra of octanol (Figure 8.2) and ethanol (Figure 8.3) in the organic solvents display a bimodal character. The bimodal character is revealed by

Gaussian fits as shown in Figure 8.2. The lack of a third component peak can be rationalized if two of the types of free O−H groups have frequencies very close to each other and thus cannot be resolved. However, it is also possible that the bimodal character

(asymmetric character to the free O−H bands) is derived from two types of solute-solvent complexes. Solute-solvent complexation is further explored in the computational studies below.

366 It is important to note that lower alcohol concentrations in the inert solvents were also examined, but the free O−H band was below the detection limit of the instrument. In addition, experiments were conducted using anhydrous ethanol and benzene and cyclohexane dried with anhydrous magnesium sulfate, and the trends in the Raman spectra were reproduced. Polarized Raman studies were also completed, and the peak shifts were reproduced with the majority of the intensity from the isotropic component

(polarized parallel to the electric field vector of the incident vertically polarized laser beam).

8.3.3. Computational spectra of ethanol/solvent complexes. To further understand the

spectral character of the O−H band of the alcohol with respect to changing organic

solvent, minima for the 1:1 complexes of ethanol and a coordinated solvent molecule were determined by optimizing a wide range of starting geometries. For the 1:1 complexes of benzene and ethanol, geometries 8.1-8.3 were identified as minima by both the HF and MP2 methods (Figure 8.4). For the other ethanol complexes (with cyclohexane and with carbon tetrachloride), trends were less consistent. For the ethanol/cyclohexane complexes (Figure 8.5), the HF computations determined three geometries: 8.4-8.6. Of these minima, MP2 only returned 8.4 and 8.6. In the case of the

ethanol/carbon tetrachloride complexes (Figure 8.6), HF and MP2 each determined three

minima. HF identified 8.7-8.9; MP2 duplicated 8.8-8.9 and also provided 8.10, in which the hydroxyl group of ethanol faces carbon tetrachloride perpendicularly, rather than 8.7.

H-bonding between the H of ethanol's hydroxyl group and the relevant solvent molecule

was observed in complexes 8.1-8.6 and 8.10. In complexes 8.7-8.9, it appears that the

367 primary interaction was due to van der Waal's forces. Analyses at the NPA level, using the B3LYP/6-31+G**//MP2/6-31G* level of theory, revealed no significant electrostatic component to any of the interactions.

2. 753 2 . 845 3. 298 2 . 484 2. 665 2. 405

8.1 8.2 8.3

Figure 8.4. Ethanol:benzene complexes as calculated at the HF/6-31G* (top) and MP2/6- 31G* (bottom) levels of theory. Distances between species are shown in angstroms.

2.962 3.437 3.553 2.612 2.699

8.4 8.5 8.6

368

Figure 8.5. Ethanol:cyclohexane complexes as calculated at the HF/6-31G* (top) and MP2/6-31G* (bottom) levels of theory. Distances between species are shown in angstroms. The geometry of 8.5 did not converge at the MP2/6-31G* level, so only the HF/6-31G* interspecies distance is shown.

8.7 4.276 8.8 4.294 3.337

3.116 3.150 8.9 8.10 2.878

Figure 8.6. Ethanol:CCl4 complexes as calculated at the HF/6-31G* (top) and MP2/6- 31G* (bottom) levels of theory. Distances between CCl4 and EtOH are shown in angstroms. The geometry of 8.7 did not converge at the MP2/6-31G* level, so only the HF/6-31G* distance is shown; the geometry of 8.10 did not converge at the HF/6-31G* level so only the MP2/6-31G* distance is shown.

The simplest way to compare experimental results to theoretical predictions is through an examination of their respective Raman spectra (Figure 8.7). These simulated

Raman spectra were generated from the calculated Raman frequencies and intensities

(MP2/6-31G*) and the B3LYP/6-31+G**//MP2/6-31G* derived Boltzmann population

369 a) b)

ethanol ethanol

ethanol/benzene complexes ethanol/benzene complexes

ethanol/cyclohexane complexes ethanol/cyclohexane complexes

ethanol/CCl4 complexes ethanol/CCl4 complexes

3520 3540 3560 3580 3600 3520 3540 3560 3580 3600 Calculated Raman Shift (cm-1) Calculated Raman Shift (cm-1) of each complex, calculated from the relative H0 (i.e. bottom-of-the-well energy + scaled

ZPE) value of each complex.

Figure 8.7: Gas-phase (a) and PCM (b) calculated Raman spectra, generated via MP2/6- 31G* calculations and Boltzmann weighting factors from H0 values (from B3LYP/6- 31+G**//MP2/6-31G* energies).

Several energy expressions were explored in addition to H0: the bottom-of-the-well energies, as well as H298 and G298 values. The Boltzmann weighting factors provided via each method are compiled in Table 8.1. The best match to the experimental spectra was seen with the H0 energies; it is presumed that this 0 K value bypasses any error introduced via the calculated thermal and entropic corrections to the enthalpy and free

370 energy, as the smallest, and least confident, vibrational frequencies make the largest contribution to the thermal and entropic corrections (for comparison, spectra calculated with all energies and compiled Boltzmann weighting factors obtained using all quantities are included in the supporting information).

EtOH: Benzene EtOH: Cyclohexane EtOH: CCl4 Gas-phase 8.1 8.2 8.3 8.4 8.6 8.8 8.9 8.10

EBW 0.33 0.48 0.19 0.63 0.37 0.04 0.90 0.06

H0 0.33 0.48 0.20 0.63 0.37 0.04 0.90 0.06

H298 0.34 0.48 0.18 0.63 0.37 0.04 0.89 0.06

G298 0.61 0.26 0.13 0.90 0.11 0.11 0.86 0.02 PCM

EBW 0.25 0.43 0.32 0.60 0.40 0.08 0.90 0.02

H0 0.25 0.43 0.32 0.60 0.40 0.08 0.90 0.02

H298 0.26 0.44 0.30 0.60 0.40 0.09 0.89 0.02

G298 0.51 0.26 0.23 0.88 0.12 0.21 0.79 0.01

Table 8.1. Boltzmann weighting factors, as calculated from various energy expressions, using B3LYP/6-31+G**//MP2/6-31G* single-point energies. The top set of values refers to gas-phase energies; the bottom, to PCM energies.

Because MP2 frequency calculations are generally considered more reliable than

HF,39a the spectra generated via the MP2 method will be the major topics of focus.39b The

371 O−H stretch of the ethanol:benzene spectrum somewhat overlaps with that of ethanol

alone. There is a distinct shoulder to the complex’s peak, due to 8.1’s unique O−H

stretch, at a considerably higher energy than that of 8.2 or 8.3. [For the remainder of this

discussion, we will focus on ethanol/benzene complexes 8.2 and 8.3 since, as reflected in

the relative areas of the peak and its shoulder, complexes 8.2 and 8.3 (main peak) have a

larger Boltzmann weighting than complex 8.1 (shoulder).] For the ethanol:cyclohexane

spectrum, no such duality was seen; however, the conformers’ spectra were slightly red-

shifted compared to that of ethanol alone. This was comparable to the spectrum seen for

ethanol:carbon tetrachloride, which also showed a single, red-shifted peak.

The MP2 calculations predicted a red shift for all three complexes. As seen in the

experimental data, only the benzene complex underwent such a substantial shift. The

cyclohexane and CCl4 complexes experimentally demonstrated narrower O−H peaks than ethanol alone, which again was predicted by the MP2 calculations. Also, the MP2- generated spectrum of the benzene complex saw a distinct shoulder, as discussed above, which was duplicated in the broad slope of the experimental O−H peak. (The HF calculations show no substantial change in either the cyclohexane or carbon tetrachloride complexes and actually predict a blue shift for the benzene complexes. Overall, the MP2 calculations more readily compare to the experimental spectra.)

The disparity in the quantitative value between the experimental and theoretical

spectral peaks might be rationalized since the calculated spectrum reflects the gas-phase

energies and vibrational frequencies of the different species. However, we did perform

single-point polarizable continuum model (PCM) energy calculations at the B3LYP/6-

372 31+G** level of theory, in order to compute a solution-level Boltzmann weighting, and

no significant spectral changes resulted (Figure 8.7).54

What is evident from these spectra is that the qualitative trends are largely

duplicated between the experimental and calculated spectra. This is most clear in the

benzene spectra, where both the red shift (with the sole exception of complex 1) and the

distinctive bimodal nature of the O−H stretch are duplicated between theory and experiment. In these spectra, the bimodal character is due to multiple isomers contributing to the Boltzmann distribution. With the more straightforward CCl4 and cyclohexane complexes, the red shift of each peak is still replicated.

8.3.4. Rationalization of experimental findings. For most geometries, regardless of method used, complexation with a solvent molecule increased the O−H bond length to a extent (Table 8.2). These results can be compared circuitously to the experimental findings. The cyclohexane and CCl4 complexes revealed minor changes in the O−H bond

length compared to ethanol; correspondingly, the O−H stretches are of similar energy and

their spectra are comparable to ethanol alone. All three of the benzene complexes’ O−H

bonds increased in length, and the corresponding vibrations decreased substantially in

energy. Thus, the considerable red shift seen for the benzene/ethanol complexes is

attributed in some measure to a lengthening of the O−H bond.

373

Scaled O−H Raman activity Bond length frequency Geometry (Å 4/AMU) (Å) (cm-1)

Ethanol 0.971 3567 97

Benzene: 8.1 0.972 3578 169 Ethanol 8.2 0.973 3545 144

8.3 0.972 3550 80

Cyclohexane: 8.4 0.971 3554 70

Ethanol 8.6 0.972 3554 65

CCl4: Ethanol 8.8 0.971 3562 98 8.9 0.972 3552 79

8.10 0.972 3568 104

Table 8.2. O−H bond lengths (Å) and stretching frequencies (cm-1) for each distinct complex, as calculated and scaled (factor of 0.9427) at the MP2/6-31G* level of theory. Raman activities (Å 4/AMU) are included and reported as unnormalized values.

8.3.5 Survey of computational methods and basis sets. In addition to supplementing the

experimental work, computational methods were also used to investigate the efficacy of various methods and basis sets. The relative energies within each set of complexes were explored using a variety of energy expressions and basis sets, including the B3LYP

374 method with the flexible 6-311+G(3df,2p) basis set. The bottom-of-the-well electronic

energies (EBW) were compiled, as were the enthalpies at 0 K, the enthalpies at 298 K, and

the free energies at 298 K, and expressed as relative energies (Table 8.3). Additionally,

the energies of the complexes relative to the energies of the complexes’ individual parts

(i.e. infinitely separated ethanol and solvent molecules) are included in Supporting

Information.

The most accurate energies (B3LYP/6-311+G(3df,2p)//MP2/6-31G*) agree well,

reproducing trends among “types” of energy: the most stable benzene complex, 8.2, is the

most energetically favorable of the three complexes in all cases except that of ΔG298

(which is a common exception in the energy trends reported here, reasserting the possibility that inaccuracy might arise in the complexes’ respective entropic corrections), where 8.1 is slightly more stable. These differences in magnitude and sign are also seen for the cyclohexane and CCl4 complexes, albeit to different extents than with the benzene

species; the general pattern of ΔEBW ≈ ΔH0 < ΔH298 < ΔG298 is seen regardless of solvent.

HF predicts more negative values for the complexation energies than does MP2

(Supporting Information), and a few slight disparities are observed between their trends;

the B3LYP/6-311+G(3df,2p)//HF/6-31G* energies predict 8.1 to be the most stable

ethanol:benzene complex regardless of quantity of interest, while in cyclohexane and

CCl4, the same overall minima are seen as with MP2 (8.4 and 8.9, respectively), except

for slight disparities in the ΔG298 values.

375

B3LYP/6-311+G(3df,2p)//MP2/6-31G* B3LYP/6-31+G**//MP2/6-31G*

ΔEBW ΔH0 ΔH298 ΔG298 ΔEBW ΔH0 ΔH298 ΔG298

EtOH: C6H6 8.1 0.12 0.13 0.10 -0.61 0.22 0.24 1.14 -0.50 8.2 0 0 0 0 0 0 0 0

8.3 0.71 0.70 0.73 0.58 0.53 0.54 1.50 0.41

EtOH: 8.4 0 0 0 0 0 0 0 0 C6H12 8.6 0.18 0.16 0.18 1.14 0.31 0.31 0.31 1.27

EtOH: CCl4 8.8 1.49 1.51 1.44 0.86 1.84 1.85 1.78 1.20 8.9 0 0 0 0 0 0 0 0

8.10 1.28 1.28 1.27 1.81 1.59 1.59 1.57 2.11

Table 8.3: Relative energies (in kcal/mol) compiled using two distinct basis sets. Energies are represented relative to that of the lowest-energy complex within each series.

Generally, increasing the size of the basis set for the single-point energies made the complexation energy more positive for a given species but did not disrupt the relative energies of the complexes. There are disparities between HF and MP2 single-point energies (in the most glaring example, 8.1 is favored exclusively by the single-point energy calculations with the HF geometries while 8.2 is favored by the MP2 geometries), but within methods, the trends between single-point energies are consistent. Thus, the

376 B3LYP/6-31+G** level provides a less computationally expensive, comparably accurate alternative to the larger and more costly 6-311+G(3df,2p) basis set. Similar findings have been noted in other systems with the potential for hydrogen bonding.7

8.4 Conclusions

The interactions of the free O−H bonds in n-octanol and ethanol with the organic solvents benzene, carbon tetrachloride, and cyclohexane were examined using Raman spectroscopy. The Raman spectra of n-octanol and ethanol complexes reveal that although cyclohexane and carbon tetrachloride as solvents have a small effect on the alcohol free O−H peak, benzene as a solvent significantly red-shifts the predominant free

O−H peak. Computational analyses matched these solvent effects and supported the experimental findings: calculated spectra generated via MP2/6-31G* frequencies, which were Boltzmann-weighted using both gas-phase and solution-phase (PCM) single-point energies, were consistent with the experimental spectra. The observed red shift in the

Raman spectra of the alcohol/benzene complexes is attributed to a lengthening of the

O−H bond from the O−H interaction with the delocalized electronic structure of benzene.

The bimodal character of the free O−H peak of the alcohol/benzene mixtures is consistent with the calculated minima for two distinctly different types of alcohol/benzene complexes.

377 References for Chapter 8

(1) Kanno, H.; Honshoh, M.; Yoshimura, Y. J. Solution Chem. 2000, 29, 1007-1016.

(2) Sassi, P.; Morresi, A.; Paolantoni, M.; Cataliotti, R. S. J. Mol. Liq. 2002, 96-97, 363-

377.

(3) Jeffrey, G. A.; Saenger, W. Hydrogen Bonding in Biological Structures. Springer-

Verlag: Berlin, 1991.

(4) Whitesides, G. M.; Mathias, J. P.; Seto, C. T. Science 1991, 254, 1312-1319.

(5) Desiraju, G. R. Nature 2001, 412, 397-400.

(6) Conn, M. M.; Rebek, J., Jr. Chem. Rev. 1997, 97, 1647-1668.

(7) Callam, C. S.; Singer, S. J.; Lowary, T. L.; Hadad, C. M. J. Am. Chem. Soc. 2001,

123, 11743-11754.

(8) Di Laudo, M.; Whittleton, S. R.; Wetmore, S. D. J. Phys. Chem. A 2003, 107, 10406-

10413.

(9) Ma, G.; Allen, H. C. J. Phys. Chem. B 2003, 107, 6343-6349.

(10) Liu, D.; Ma, G.; Levering, L. M.; Allen, H. C. J. Phys. Chem. B 2004, 108, 2252-

2260.

(11) Zheng, J.; Kwak, K.; Asbury, J.; Chen, X.; Piletic, I.R.; Fayer, M. D. Science 2005,

309, 1338-1343.

378 (12) Lalanne, P.; Andanson, J. M.; Soetens, J.-C.; Tassaing, T.; Danten, Y.; Besnard, M.

J. Phys. Chem. A 2004, 108, 3902-3909.

(13) Zwier, T. S. Annu. Rev. Phys. Chem. 1996, 47, 205-241.

(14) Provencal, R. A.; Casaes, R. N.; Roth, K.; Paul, J. B.; Chapo, C. N.; Saykally, R. J.;

Tschumper, G. S.; Schaefer III, H. F. J. Phys. Chem. A 2000, 104, 1423-1429.

(15) Augspurger, J. D.; Dykstra, C. E.; Zwier, T. S. J. Phys. Chem. 1993, 97, 980-984.

(16) Suzuki, S.; Green, P. G.; Bumgarner, R. E.; Dasgupta, S.; Goddard III, W. A.; Blake,

G. A. Science 1992, 257, 942-945.

(17) Augspurger, J. D.; Dykstra, C. E.; Zwier, T. S. J. Phys. Chem. 1992, 96, 7252-7257.

(18) Garrett, A. W.; Severance, D. L.; Zwier, T. S. J. Chem. Phys. 1992, 96, 7245-7258.

(19) Garrett, A. W.; Zwier, T. S. J. Chem. Phys. 1992, 96, 7259-7267.

(20) Pribble, R. N.; Hagemeister, F. C.; Zwier, T. S. J. Chem. Phys. 1997, 106.

(21) Gruenloh, C. J.; Florio, G. M.; Carney, J. R.; Hagemeister, F. C.; Zwier, T. S. J.

Phys. Chem. A 1999, 103, 496-502.

(22) Veldhuizen, R.; de Leeuw, S. W. J. Chem. Phys. 1996, 105, 2828-2836.

(23) Akiyama, I.; Ogawa, M.; Takase, K.; Takamuku, T.; Yamaguchi, T.; Ohtori, N. J.

Sol. Chem. 2004, 33, 797-809.

(24) Bogdanov, B.; McMahon, T. B. J. Phys. Chem. A 2000, 104, 7871-7880.

379 (25) Cramer, C.; Truhlar, D. Chem. Rev. , 1999, 99, 2161-2200.

(26) Tomasi, J. Chem. Rev., 1994, 94, 2027-2094.

(27) Bauer, E.; Magat, M. J. Phys. Radium 1938, 9, 319-330.

(28) Onsager, L. J. Am. Chem. Soc. 1936, 58, 1486-1493.

(29) Beak, P. Acc. Chem. Res. 1977, 10, 186-192.

(30) Chandrasekhar, J.; Smith, S. F.; Jorgensen, W. L. J. Am. Chem. Soc. 1985, 107,

154-163.

(31) Gong, X.; Xiao, H. J. Mol. Struct. (THEOCHEM) 1998, 453, 141-147.

(32) Rivail, J. L.; Rinaldi, D.; Dillet, V. Mol. Phys. 1996, 89, 1521-1529.

(33) Ramirez, F. J.; Tunon, I.; Silla, E. J. Phys. Chem. B 1998, 102, 6290-6298.

(34) Ilich, P.; Hemann, C. F.; Hille, R. J. Phys. Chem. B 1997, 101, 10923-10938.

(35) Orozco, M.; Luque, F. J. Chem. Rev. 2000, 100, 4187-4226.

(36) Dimitrova, Y. J. Mol. Struct. (THEOCHEM) 1997, 391, 251-257.

(37) McCreery, R. L. Raman Spectroscopy for Chemical Analysis. John Wiley & Sons:

New York, 2000.

(38) Frisch, M. J. T., G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J.

R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.;

Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.;

380 Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa,

J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J.

E.; Hratchian, H. P.; Cross, J. B.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R.

E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.;

Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.;

Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.;

Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.;

Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.;

Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.;

Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A.; . Gaussian 03, Revision B.03. Gaussian, Inc.: Pittsburgh, PA, 2003.

(39) a.) Hehre, W. J.; Radom, L.; Schleyer, P. v. R.; Pople, J. A. Ab Initio Molecular

Orbital Theory. John Wiley & Sons: New York, 1986; b.)Additional levels of theory were

considered, and QCISD calculations were run on the benzene:ethanol complexes,

confirming the geometries identified by HF and MP2. However, these calculations were

not explored further, due to their high computational cost.

(40) Johnson, B.G.; Gill, P.M.W.; Pople, J.A. J. Chem. Phys. 1993, 98, 5612-5626.

(41) Lee, C.; Yang, W.; Parr, R.G.; Phys. Rev. B 1988, 37, 785-789.

(42) Stephens, P.J.; Devlin, F.J.; Frisch, M.J. J. Phys. Chem. 1994, 98, 11623-11627.

(43) Becke, A.D. J. Chem. Phys. 1993, 98, 5648-5652.

381 (44) Barckholtz, C.; Barckholtz, T. A.; Hadad, C. M. J. Am. Chem. Soc. 1999, 121, 491-

500.

(45) (a) Pople, J. A.; Scott, A. P.; Wong, M. W.; Radom, L. Isr. J. Chem. 1993, 33, 345-

350. (b) Scott, A. and Radom, l.; J. Phys. Chem. 1996, 100, 16502-16513.

(46) Cossi, M.; Barone, V.; Cammi, R.; Tomasi, J. Chem. Phys. Lett. 1996, 255, 327-

335.

(47) Barone, V.; Cossi, M.; Tomasi, J. J. Chem. Phys. 1997, 107, 3210-3221.

(48) Barone, V.; Cossi, M.; Tomasi, J. J. Comput. Chem. 1998, 19, 404-417.

(49) Cossi, M.; Barone, V. J. Chem. Phys. 1998, 109, 6246-6254

(50) Mammone, J. F.; Sharma, S. K. J. Phys. Chem. 1980, 84, 3130-3134.

(51) Daimay, L.-V.; Colthup, N. B.; Fateley, W. G.; Grasselli, J. G. The Handbook of

Infrared and Raman Characteristic Frequencies of Organic Molecules. Academic Press:

Boston, 1991.

(52) Graener, H.; Ye, T. Q.; Laubereau, A. J. Chem. Phys. 1989, 90, 3413-3416.

(53) Woutersen, S.; Emmerichs, U.; Bakker, H. J. J. Chem. Phys. 1997, 107, 1483-1490.

(54) Further work might incorporate a solution-phase optimization, so as to improve

(change) the calculated vibrational frequencies and intensities, which could include

consideration of solute-solvent interaction or the solvents’ dielectric constants, to more

accurately match numerical values.

382

CHAPTER 9

THE EFFECTS OF HYDROGEN AUGMENTATION ON THE COMBUSTION BEHAVIOR OF HYDROCARBON FUELS: FUEL BLENDS AND MASTER EQUATION ANALYSES

9.1. Introduction

9.1.1. Hydrogen as fuel. A clean-burning, readily available alternative fuel has long been a target of environmental scientists, and public demand has correspondingly increased in recent years, especially with rising gasoline prices. Hydrogen is one likely candidate; it can be generated in its entirety from a renewable source, water (although the generation process involves much energy). As such, it has been a target of interest for scientists for several decades.1,2

Hydrogen demonstrates improved combustion properties relative to fossil fuels in many aspects, from fuel efficiency and power generation to advantageous environmental emissions. Compared to typical hydrocarbon fuels such as gasoline, hydrogen exhibits higher flame propagation rates, flame speeds, and shorter combustion times, which lead to increased engine efficiencies.3 Additionally, it has a wide range of flammability limits, and therefore, it can be used over a wide range of concentrations. In particular, since hydrogen has a low lean operational limit, it allows fuel-lean operation (Φ ~ 0.2) of

383 an engine,4 which in turn leads to greater engine control relative to hydrocarbon fuels, as

well as increased fuel economy, and fewer emissions.5 This final aspect is perhaps the

most-cited benefit of hydrogen as a fuel. It is obvious that hydrogen would avoid

production of carbon monoxide, carbon dioxide, and hydrocarbon emissions;

correspondingly, it also cuts down on particulate emissions and smoke, themselves

byproducts of hydrocarbon emission. Finally, hydrogen also minimizes (or eliminates)

SOx and NOx emissions; the latter is an unusual case. Hydrogen combustion itself may

result in substantial NOx production; however, since hydrogen can be used at such fuel- lean conditions, hydrogen as a fuel counters this trend to decrease the overall emission of

NOx. Finally, besides its improved combustion characteristics, hydrogen also affords considerable benefits in engine performance, such as reduced deposits and engine wear, relative to gasoline, increased oil life, and higher compression ratios.6 However, relative to gasoline as a pure fuel, hydrogen creates other challenges in fuel input and engine design due to its flammability and pre-ignition at lower temperatures.4-6 These properties and a few others are summarized in Table 9.1.

The combustion mechanism of hydrogen is relatively straightforward. Several

mechanisms have been developed for the subsets of reaction that dictate the oxidation of

this simple fuel; Table 9.2 summarizes some of the key processes, rates, and

thermodynamic data, as well as the researchers7,,8 9 who developed this information.

These data are written in Arrhenius form (Equation 9.1):

−Ea b RT kf = AT e (9.1)

384 In the Arrhenius equation, kf is the rate coefficient for the forward reaction, T is temperature, A is the pre-exponential factor; b is a measure of the temperature dependence, and Ea is the activation barrier of a reaction. Each elementary step possible for hydrogen has thus been outlined, each via its own kinetic experiment or extrapolation from similar work, to compile the entire mechanism for hydrogen combustion.

Property Hydrogen Gasoline Diesel

Fuel-air ratio 1/34.5 1/14.7 1/14.5 (kg fuel/kg air)

Lower heating 119.7 43.3 42.5 value (MJ/kg)

Auto-ignition 585 400 250 temperature (K)

Flammability limits 4.0−75.0 1.0−7.6 1.0−6.5 (volume %)

Flame speed (cm/s) 265−325 37−43 -

Table 9.1. Characteristics of hydrogen as compared to typical hydrocarbon fuels.4-6

385

Elementary Reaction Step A b Ea Reference

1 OH + H2 = H + H2O 2.14 E +08 1.52 3.449 a

2 O + OH = O2 + H 2.02 E +14 -0.4 0 a

3 O + H2 = OH + H 5.06 E +04 2.67 6.29 a

4 H + O2 + M = HO2 + M 4.52 E +13 0 0 a

5 OH + HO2 = H2O + O2 2.13 E +28 0 3.5 b

6 H + HO2 = OH + OH 1.50 E +14 0 1 a

7 H + HO2 = H2 + O2 6.63 E +13 0 2.126 c

8 H + HO2 = O + H2O 3.01 E +13 0 1.721 a

9 O + HO2 = O2 + OH 3.25 E +13 0 0 a

10 2OH = O + H2O 3.57 E +04 2.4 -2.112 a

11 H + H + M = H2 + M 1.00 E +18 -1 0 a

12 H + OH + M = H2O + M 2.21 E +22 -2 0 a

13 H + O + M = OH + M 4.71 E +18 -1 0 a

14 O + O + M = O2 + M 1.89 E +13 0 -1.788 a

15 HO2 + HO2 = H2O2 + O2 4.20 E +14 0 11.982 a

16 OH + OH + M = H2O2 + M 1.24 E +14 -0.37 0 a

17 H2O2 + H = HO2 + H2 1.98 E +06 2 2.435 a

18 H2O2 + H = OH + H2O 3.07 E +13 0 4.217 a

19 H2O2 + O = OH + HO2 9.55 E +06 2 3.97 a

20 H2O2 + OH = H2O + HO2 2.40 E 00 4.042 -2.162 a (a) Ref. 7 (b) Ref. 8 (c) Ref. 9

Table 9.2. Arrhenius coefficients for the combustion of hydrogen. Units in moles, cm3, seconds, and K (A, b); kcal/mol (Ea).

386

This combustion mechanism clarifies many benefits seen with hydrogen combustion relative to hydrocarbon combustion. The steps included in Table 9.2 occur quickly (low activation barriers); the chain-branching steps are generally thermoneutral overall. In hydrocarbon combustion, the chain-branching steps are endothermic and invoke higher reaction barriers. These kinetic effects instigate significant differences in overall oxidation characteristics of the two types of fuels.

9.1.2. Hydrogen as fuel additive. Hydrogen also exhibits several drawbacks, as a fuel in its own right. Hydrogen has less energy content than an equal volume of gasoline, and power output is severely reduced for engines fueled by hydrogen. Hydrogen engines also undergo a great deal of backfiring and uncontrolled pre-ignition,10 and high combustion temperatures and pressures are generated when engines operate at near-stoichiometric conditions.6 One method to circumvent these issues is to use hydrogen as a fuel additive with gasoline or another hydrocarbon fuel; this approach combines the benefits of hydrogen with those of hydrocarbon fuels.

9.1.2.1. Generating hydrogen. Although blending hydrogen and hydrocarbon fuels to exploit the benefits of each seems like a reasonable approach, the use of hydrogen has been limited, due to the fact that it cannot be reliably generated in an energetically economical manner. Several solutions have been put forth to bypass this problem, including storing hydrogen as a compressed gas, generating hydrogen by providing heat to metallic hydrides,11 and carrying hydrogen as a cryogenic liquid,12 but all pose

387 significant costs and drawbacks. One reasonable method for the addition of hydrogen to a given fuel mixture involves an electrolyzer, which electrochemically converts water to hydrogen and oxygen via a safe and cost-effective manner. When coupled to a gasoline engine, this device converts chemical energy generated from fuel combustion to electrical energy, which is used in turn to convert water to hydrogen and oxygen, which can then be used in augmenting fuel. Dulger and Ozcelik reported on one instance of the use of an on-board electrolyzer to generate hydrogen production and improve fuel economy; they noted that use of their electrolyzer resulted in improvements in fuel consumption ranging from 26−43%, depending on the automobile in question, and reduced exhaust emissions.13 (One ongoing concern with electrolyzers is the balance between the power

necessary to generate the electrolysis products and the power gained by the engine as a

result of hydrogen augmentation.) Another potential method of hydrogen generation

involves reforming exhaust gas, thus decreasing emissions, such as smoke and NOx, while generating H2 for fuel augmentation. This method has been reported particularly

14,15 for use with diesel engines, by Tsolakis et al.

9.1.2.2. Methane/H2. Uyker et al. have examined the effect of addition of electrolysis

products on methane/air premixed laminar combustion, using the CHEMKIN simulation

package.16 They saw that hydrogen augmentation (10−20%) did have a small beneficial

effect on flame speed and on lean flammability limit properties. Moreover, when both

the electrolysis products (hydrogen and oxygen) were considered, the addition of 10%

hydrogen and the associated oxygen matched the effects seen from addition of 20%

hydrogen alone. NOx emissions were predicted to increase when oxygen addition was

388 considered; carbon monoxide emissions were predicted to decrease when hydrogen was

present in place of hydrocarbon fuels.

9.1.2.3. Gasoline/H2. In a gasoline engine, a homogeneous environment where

combustion rate is primarily dictated by flame speed, hydrogen augmentation has been shown to have several effects in enhancing overall combustion. Generally speaking, combustion effects are highly sensitive to several variables, including the amount of hydrogen augmentation, the overall equivalence ratio, and engine conditions.

In many of these studies, benefits of augmentation “peak” at particular

percentages of hydrogen. Hacochen and Sher have demonstrated that a nearly 15%

increase in fuel economy is possible with <6% hydrogen enrichment, but also noted that beyond this point, the benefits decrease.17 Al-Janabi and Al-Baghdadi examined the

18 effect of H2 augmentation in gasoline on engine performance and fuel emissions; they

saw a maximum improvement in engine thermal efficiency at 8% H2 blending and a maximum reduction in fuel consumption at 6% H2 blending. In terms of emissions, a

10% increase in H2 decreased carbon monoxide concentrations by 74% but doubled NO

concentrations (although this could be alleviated by varying equivalence ratios).

In other gasoline/H2 studies, Sobiesiak et al. reported on the effect of H2/O2 additives on a premixed iso-octane/air flame,19 noting increases in flame speed and

adiabatic flame temperature. They also assigned the resultant decrease in carbon

monoxide emissions to increased concentrations of hydroxyl (HO•) radical. Henshaw et

al. completed a related study, exploring the specific effects of H2 + O2 augmentation (as a

result of water electrolysis) on a gasoline-fueled engine.20 They saw that hydrogen

389 augmentation alone caused essentially the same effects as hydrogen and oxygen;

moreover, as the equivalence ratio of the fuel mixture neared 1.0, the beneficial effects on

flame speed and other variables tapered off. Finally, they observed that the power

necessary to generate the electrolysis products of hydrogen and oxygen outweighed the

power gained by the engine as a result of hydrogen augmentation.

9.1.2.3. Diesel/H2. In the diesel engine, since H2 is present as a gas, while diesel fuel exists as liquid droplets, combustion is a heterogeneous scenario. The two types of fuel cannot interact except in the vicinity of the diesel droplets. Given the disparate auto- ignition temperatures of diesel and hydrogen, as well as their heterogeneity, the two fuels burn independently. Thus, the chemical effects of hydrogen augmentation differ from the gasoline engine described above: the low ignition energy, high flame speed, and wide mixture range are not as relevant to diesel combustion.

Naber and Siebers used a constant-volume combustion vessel with simulated

diesel conditions to study the effects of hydrogen augmentation.21 They noted that

hydrogen can combust under diesel conditions; its ignition delay exhibited an Arrhenius

dependence on temperature, but was essentially independent of oxygen concentration and

pressure. Kumar et al. studied the effects of hydrogen blending on both diesel fuels and

vegetable oils (biodiesel);22 they saw that vegetable oils alone demonstrated reduced

efficiency and increased smoke relative to diesel fuel, but hydrogen augmentation

improved these effects. For both diesel and biodiesel, hydrogen augmentation reduced

smoke, hydrocarbon, and carbon monoxide emissions, as well as overall combustion

390 duration. NOx emissions increased, as did the ignition delay and the rate of pressure

increase. They noted that these effects peaked at 5% hydrogen augmentation.

9.1.3. Master equation methods and fuel behavior simulations. Once a comprehensive

oxidation mechanism is delineated for a given fuel, these data can be used with master

equation methods to model fuel performance in various reactors. As shown with

hydrogen (Table 9.2), several reaction steps are possible even for a small molecule:

mechanism development demands a significant amount of time and effort. Methane

combustion, for instance, requires 325 steps to be completely described.23 Hydrocarbon

fuels, such as n-heptane24 and iso-octane,25 are components of the Primary Reference

Fuel (PRF) mechanism, developed by Curran et al., which is commonly used in

understanding gasoline oxidation.26 n-Hexadecane shows promise for understanding

diesel oxidation.27 Also notable is the fact that mechanism development is a cooperative

process: hydrogen combustion is an important subset of each subsequently developed

mechanism (methane, ethane, etc.), so that these reactions have implications even when hydrogen is not the actual fuel under consideration.

A master equation analysis requires three types of data files: kinetics, comprised

by the Arrhenius parameters (A, b, and Ea) for each elementary reaction step;

thermodynamics, consisting of coefficients to polynomial fits to the specific heats,

enthalpies, and entropies of each species involved in the mechanism; and transport,

including data on polarizability, dipole moments, and Lennard-Jones properties. A

master equation method uses these files to generate input parameters and a series of

differential equations that are ultimately solved to give the concentrations of the

391 reactants, intermediates, and products spatially and temporally (Figure 9.1). This information can then be extended to determine other properties of a given reaction mixture, depending on the reactor model of interest.

KINETIC MECHANISM

THERMODYNAMIC RESULTS DATA REACTOR MODEL

TRANSPORT DATA

Figure 9.1. Generalized depiction of the progress of a master equation method.

It may be instructive to discuss these steps in the context of a specific master equation method. CHEMKIN 4.1 was used to complete the work discussed in this chapter. Generated by researchers at Sandia National Labs28 and currently maintained by

Reaction Design,29 this software package can model a wide variety of experimental settings, via geometric approximations. This work used the PREMIX module to model a one-dimensional flame; a premixed flame is one in which the fuel and oxidizer are completely mixed prior to ignition. After generating the relevant input files, CHEMKIN

392 implements the TWOPNT solver30 in addressing the differential equations; this method

modifies a hybrid, damped Newton algorithm,31 along with time stepping; for our

purposes, this means that the initial guess of species concentrations does not have to be

exact. Once the differential equations are solved, CHEMKIN’s post-processing software

generates profiles of temperature, flame speed, species concentration, emissions, etc., as

functions of distance from the ignition point.

9.2. Computational Methods. The master equation method CHEMKIN 4.1 was used to simulate the properties of various fuel blends. For the majority of these cases, the primary reference fuels (PRF) constituted the main species of interest. In other runs, n- heptane was also considered. Curran et al. have generated comprehensive kinetics, thermodynamics, and transport files for both the PRF25-26 and n-heptane mechanisms.

These initial files were then altered to account for varying amounts of H2 augmentation;

with the PRF mechanism, the H2 augmentation varied from 0−10 %; with the n-heptane mechanism, augmentation varied from 0−50 %.

The freely-propagating flame model in PREMIX was used to model combustion

in these different fuel mixtures. In this module, ignition and combustion were modeled

over a distance of 0.3 cm. An initial temperature profile was used as part of the input file

and set ignition at 0.05 cm. The PRF mechanism in particular was complicated; the time

limits and the single-processor limitation of the calculations were such that only the

initial 0.3 cm run could be completed within the 300-hour time limit that constituted the

maximum time provided by the Ohio Supercomputer Center. For the smaller

393 mechanisms, it was possible to refine the calculations and expand out to a larger distance

(5−10 cm), in the allotted time.

In all other respects, the default values given with the CHEMKIN module were

used, using a pressure of 1.0 atm and an initial flow rate of 0.04 g/(cm2.s). An initial grid

of six points was set for solving the differential equations, and additional points were

added by the program as necessary. Starting equivalence ratios are noted throughout the

Results and Discussion (the equivalence ratio is a measure of the ratio of fuel-to-oxidizer

in a given mixture, relative to the stochiometric ratio of fuel-to-oxidizer necessary for

combustion). Following completion of the calculations, the Post-Processor in

CHEMKIN was used to convert the results into a graphical format.

9.3. Results and Discussion

9.3.1. Primary Reference Fuels and Hydrogen. Runs were completed for 0, 1, 2, 4, 5,

and 10 % H2 augmentation relative to the starting concentrations of the primary reference

fuels (n-heptane and iso-octane). Graphs were generated to account for variation in

flame speed (Figures 9.2), hydrocarbon fuel consumption (Figure 9.3), H2 mole fraction

(Figure 9.4), carbon monoxide emissions (Figure 9.5), unburned fuel fractions (Figure

9.6), and net heat production (Figure 9.7). Flame speed is commonly used as a metric for flame efficiency and is generated by the Post-Processor, while the information on fuel consumption, H2 amount, and unburned fuel fractions can be generated directly from the

CHEMKIN solution file. Likely emissions are identified over the course of a CHEMKIN

solution; the software identifies reactions implicated in the formation of NO, NO2, total

394 NOx, and CO, and generates a mole fraction profile of any relevant emissions; in this case, carbon monoxide was seen to form at an appreciable rate.

Figure 9.2. Variation in flame speed with hydrogen augmentation to PRF combustion mechanism, as predicted by CHEMKIN 4.1. Entire distance, top; expanded view of final flame speed, bottom. 0% H2 augmentation denoted by solid diamonds 1% denoted by squares, 2% denoted by triangles; 4% denoted by x; 5% denoted by asterisks; 10% denoted by circles.

395

Figure 9.3. Variation in hydrocarbon fuel consumption (top, iso-octane; bottom, n- heptane) with H2 augmentation, as predicted by PRF mechanism. 0% H2 augmentation denoted by solid diamonds 1% denoted by squares, 2% denoted by triangles; 4% denoted by x; 5% denoted by asterisks; 10% denoted by circles.

396

Figure 9.4. Variation in H2 mole fraction as predicted by PRF mechanism, as a function of H2 augmentation (top) and of temperature change (bottom). 0% H2 augmentation denoted by solid diamonds 1% denoted by squares, 2% denoted by triangles; 4% denoted by x; 5% denoted by asterisks; 10% denoted by circles.

397

Figure 9.5. Variation in carbon monoxide emissions with H2 augmentation to PRF mechanism. 0% H2 augmentation denoted by solid diamonds 1% denoted by squares, 2% denoted by triangles; 4% denoted by x; 5% denoted by asterisks; 10% denoted by circles.

Figure 9.6. Variation in unburned fuel fraction with H2 augmentation to PRF mechanism. 0% H2 augmentation denoted by solid diamonds 1% denoted by squares, 2% denoted by triangles; 4% denoted by x; 5% denoted by asterisks; 10% denoted by circles.

398

Figure 9.7. Variation in heat production with H2 augmentation to PRF mechanism. 0% H2 augmentation denoted by solid diamonds 1% denoted by squares, 2% denoted by triangles; 4% denoted by x; 5% denoted by asterisks; 10% denoted by circles.

The experimental flame speed range for the PRF mechanism was shown to be 20-

40 cm/s over an equivalence ratio range of 0.7−1.4, as obtained by Huang et al.32; the

CHEMKIN-calculated values were substantially higher than these experimental values.

Subsequent work was undertaken to explore this discrepancy with a smaller mechanism.

The qualitative trends can be discussed. Overall, hydrogen augmentation does increase flame speed and fuel consumption, as well as reducing emissions; however, these effects are small, generally on the order of 1−2% of the total quantity of interest, at most.

Additionally, the trends for hydrogen augmentation of the PRF mechanism are non- linear, as shown in the case of flame speed (Figures 9.8 and 9.9).

399

Figure 9.8. Non-linear variation in flame speed at a fixed point (0.15 cm), for H2 augmentation of PRF mechanism.

Picking a fixed distance from the spark ignition and monitoring the flame speed at that distance, it was possible to determine flame speed as a function of H2 augmentation

(Figure 9.8). The 4% H2-augmentation run demonstrates a substantially higher effect

than any other run. This effect is repeated through the other observed variables; the 4%

run consistently demonstrated the most favorable trends of any of the augmentation runs,

even to a more pronounced extent than the 10% run. This is illustrated in Figure 9.9,

which shows relative trends for flame speed, carbon monoxide emissions, heat

production, and iso-octane consumption, all on the same graph. The 4% run

demonstrates the minimum value of carbon monoxide emissions and heat production, as

well as the maximum value for flame speed and iso-octane consumption. This result

parallelled previous experimental work on gasoline/H2 mixtures, which demonstrated

non-linear effects of augmentation.17-20

400

Figure 9.9. Relative overall trends with H2 augmentation to PRF mechanism. Flame speed represented by diamond; carbon monoxide emissions by small square; heat production by open triangle; iso-octane consumption by large square.

The mole fractions of any of the species of interest in this mechanism could be generated; the CHEMKIN programs generate an enormous amount of information. For the sake of simplicity, only iso-octane, n-heptane, and H2 are shown. The profiles for iso-octane and n-heptane showed a logical progression: starting with the input value and quickly dropping, once ignition occurs. However, the hydrogen profile was unusual, demonstrating a large spike in H2 production soon after ignition, which actually led to a product H2 concentration that was considerably higher than the reactant H2 concentration.

As the purpose of H2 augmentation serves to bolster combustion, it seems that this concentration should fall dramatically as distance from the burner increases. One possible explanation for this finding might involve the equilibrium between two

401 hydrogen atoms and molecular hydrogen. Rapid H-atom recombination would complete

with H2 combustion, such that the H2 profile would not follow a consistent trend.

9.3.2. n-Heptane/H2 calculations. Given the size of the PRF mechanism (1025 species and 3937 reactions), the effect of H2 augmentation was also explored for a more tractable

mechanism, that of n-heptane (154 species and 1524 reactions). Several trends were

again monitored: flame speed (Figures 9.10-9.11), n-heptane consumption (Figure 9.12),

H2 mole fraction (Figure 9.13), carbon monoxide emissions (Figure 9.14), net heat

production (Figure 9.15), and unburned fuel fractions (Figure 9.16).

402

Figure 9.10. Variation in flame speed with H2 augmentation for n-heptane (Φ = 1.2). 0 % H2 augmentation denoted by diamond; 1% denoted by square, 2% denoted by triangle; 5% denoted by x; 10% denoted by asterisk; 20% denoted by circle; 50% denoted by cross.

403

Figure 9.11. Variation in flame speed at a fixed point (2.4 cm) for n-heptane (Φ = 1.2).

Figure 9.12. Variation in n-heptane consumption with increasing H2 augmentation. 0% H2 augmentation denoted by diamond; 1% denoted by square, 2% denoted by triangle; 5% denoted by x; 10% denoted by asterisk; 20% denoted by circle; 50% denoted by cross.

404

Figure 9.13. Variation in H2 concentration with increasing H2 augmentation. 0 % H2 augmentation denoted by diamond; 1% denoted by square, 2% denoted by triangle; 5% denoted by x; 10% denoted by asterisk; 20% denoted by circle; 50% denoted by cross.

Figure 9.14. Variation in carbon monoxide emissions with increasing H2 augmentation. 0% H2 augmentation denoted by diamond; 1% denoted by square, 2% denoted by triangle; 5% denoted by x; 10% denoted by asterisk; 20% denoted by circle; 50% denoted by cross.

405

Figure 9.15. Variation in net heat production with H2 augmentation to n-heptane. 0% H2 augmentation denoted by diamond; 1% denoted by square, 2% denoted by triangle; 5% denoted by x; 10% denoted by asterisk; 20% denoted by circle; 50% denoted by cross.

Figure 9.16. Variation in unburned fuel fraction with H2 augmentation to n-heptane. 0% H2 augmentation denoted by diamond; 1% denoted by square, 2% denoted by triangle; 5% denoted by x; 10% denoted by asterisk; 20% denoted by circle; 50% denoted by cross.

406

Figure 9.17. Relative trends associated with H2 augmentation for n-heptane mechanism (Φ = 1.2). Flame speed denoted by diamonds; heptane mole fraction denoted by squares; heat production denoted by triangles.

The n-heptane calculations took far less time, as they could be run over several

hours rather than several days. This allowed calculations with greater concentrations of

hydrogen; in fact, the effects of a 50:50 mixture of n-heptane and H2 could be explored.

As with the PRF mechanism, H2 augmentation caused minor, but discernable, effects on

all the quantities of interest. Not all trends were duplicated between the PRF and n-

heptane runs: in n-heptane, increased H2 augmentation accounted for increased flame

speed and fuel consumption. However, whereas the PRF mechanism had demonstrated

that increased H2 resulted in decreased carbon monoxide emissions and heat production,

the n-heptane mechanism showed increases in these variables as well. Also, H2 augmentation to the n-heptane mechanism resulted in relatively linear trends (Figure

407 9.17). While the n-heptane results could be more quickly obtained, it remains to be seen

if these results are actually useful in fuel chemistry predictions, since they differ in a few

ways from the comprehensive PRF mechanism.

The next set of calculations sought to obtain theoretical values that could be

usefully compared to experimental work. For n-heptane, the experimental flame speed

has been shown to be 40 cm/s with Φ = 1.0; we calculated a flame speed of 160 cm/s

with Φ = 1.2. This is not so great a disparity as was previously seen with the PRF

mechanism, but it is still a cause for concern.

In this case, due to the small mechanism size, calculations could be run at various equivalence ratios, to find which theoretical equivalence ratio yielded the most reasonable flame speed (Figure 9.18). It was seen that higher equivalence ratios yielded results that were close to the experimental values; a set of calculations was then run to explore the effects of H2 augmentation on fuel-rich n-heptane combustion. The same

trends were explored (Figures 9.19 and 9.20). The qualitative trends were duplicated between the fuel-lean and fuel-rich n-heptane runs (Figures 9.17 and 9.20), and the latter demonstrated quantitatively reasonable values; the earlier calculations were circuitously shown to have some empirical relevance. Hydrogen augmentation in the n-heptane mechanism caused a rise in flame speed, an increase in heptane consumption, and a slight increase in net heat production, regardless of equivalence ratio.

408

Figure 9.18. Variation of flame speed with equivalence ratio for n-heptane, as calculated experimentally and theoretically. Experimental values denoted with open diamonds; theoretical values shown with solid squares.

409

Figure 9.19. Linear variation in fuel-rich (Φ = 2.0) flame speed with H2 augmentation to n-heptane. 0% H2 augmentation denoted by diamond; 1% denoted by squares, 2% denoted by triangles; 4% denoted by x. Bottom: variation in flame speed at fixed distance (2.4 cm)

410

Figure 9.20. Relative trends associated with H2 augmentation for fuel-rich n-heptane mechanism (Φ = 2.0). Flame speed denoted by diamonds; heptane mole fraction denoted by squares; heat production denoted by triangles.

9.4. Conclusions

Overall, calculations were successfully run for mixtures of hydrogen and hydrocarbon fuels. The effect of hydrogen as an additive varied given the identity of the parent fuel. With the primary reference fuels (represented as a mixture of n-heptane and

iso-octane), hydrogen augmentation resulted in minor increases in flame speed, decreases in carbon monoxide emissions,33 and improvements in fuel consumption; these

conclusions are in keeping with earlier experimental work on gasoline/hydrogen fuel

mixtures. However, these effects were not pronounced, and a non-linear reactivity was

seen: the 4% H2-augmentation run demonstrated the most beneficial effects with respect

411 to all of the variables explored, despite the fact that runs were completed with higher

amounts of hydrogen content. Again, this is comparable to previous work available in

the literature,17-20 which suggests that augmentation effects are highly sensitive to the

amount of hydrogen augmentation. A marked disparity was seen between predicted and

observed flame speeds (which are easily obtained experimental values). While this may

be attributed to the sensitivity of the mechanism to hydrogen augmentation, it does not

follow that the calculation for the PRF mechanism alone (with no H2 added) would

follow the same trend, if this were the case. We attribute this effect to the equivalence

ratio used in our calculations (Φ = 1.4); this value was chosen mainly on the basis of

trial-and-error, during initial CHEMKIN calculations, since we were interested in completing our calculations to observe the qualitative effects of hydrogen augmentation.

Given the time necessary to complete our initial run, we have not yet explored the effect

of a range of equivalence ratios, which could better match experimental flame speeds.

Bolstering our supposition, concurrent with the completion of this work, Mandilas et al. published their account of a study of the effects of hydrogen augmentation on iso- octane-air flames.34 As iso-octane comprises the majority of the PRF mechanism, their

results were of particular interest. In their work, the effect of 5% H2 augmentation was

consistently examined over a wide range of equivalence ratios. They noted that this

augmentation resulted in a doubling of the flame speed at the flame-lean limit, an

increase of flame speed over the 0.2 < Φ < 1.2 range, and no increase in flame speed for

Φ > 1.2. While we did see a small increase in flame speed with hydrogen augmentation for our Φ = 1.4 calculations, it was insignificant compared to the overall flame speed.

412 This supported our hypothesis that it would be necessary to select a more fuel-lean

equivalence ratio to see more pronounced quantitative effects. Previous studies have

observed a large sensitivity of combustion performance to variables such as engine type,

operating conditions, equivalence ratio, and amount of hydrogen augmentation, and the

susceptible character of these results corroborate those findings.

The n-heptane mechanism predicted linear increases in flame speed, fuel efficiency, carbon monoxide emissions, and net heat production, with hydrogen augmentation. The nature of the hydrocarbon fuel impacts CO emissions and net heat production, as these trends differed from the PRF predictions. Since n-heptane could be modeled with a smaller mechanism, at the cost of less computational time, additional calculations were run. Even at high concentrations of hydrogen, the quantitative effects of augmentation were minor.

The theoretical predictions of flame speed were substantially higher than observed in experiment. Several calculations were run with varying equivalence ratios, for the n-heptane mechanism, to find the theoretical equivalence ratio that would predict a comparable flame speed; it was seen that a more fuel-rich calculation more reliably matched the empirical findings. This was unusual given the results seen in our and previous work with gasoline modeling, in which fuel-lean calculations were more accurate. However, given the sensitivity of these results to the mechanism used, a smaller mechanism such as n-heptane may well be too simplified to model gasoline combustion. The trends seen in the n-heptane calculations were repeated regardless of equivalence ratio; however, in this case, there were limited experimental results for comparison.

413 Overall, our work has shown that hydrogen augmentation will cause only a small increase in flame speed when used with fuel-rich PRF (gasoline) mixtures. Additionally, we have seen that 4% H2 augmentation results in the most pronounced effects on flame speed and engine efficiency, of the H2/PRF fuel blends examined. Previous work shows that hydrogen augmentation would have more pronounced effect on fuel-lean PRF mixtures. In terms of the overall value of hydrogen augmentation, these results do not suggest improvement in combustion to the extent that would support the expenditure of energy and resources necessary to generate hydrogen via conventional electrolysis.

Further calculations with CHEMKIN could optimize the equivalence ratio and engine conditions for PRF combustion that would allow for greater effects of hydrogen augmentation.

414 References for Chapter 9

1 King, R. O.; Rand, M. Canad. J. Technol. 1955, 33, 445-469.

2 Erren, R. A.; Campbell, W. H. J. Inst. Fuel 1933, 6, 277-290.

3 Karim, G. A. Int. J. Hydrogen Energy 2003, 28, 569-577.

4 Bade, S.; Karim, G. A.; Wierzba, I. “Examination of operational limits in gas-fueled spark ignition engines.” SAE Paper #2000-01-1944- SP-1549.

5 Das, L. M. Int. J. Hydrogen Energy 1990, 15, 833.

6 Yamin, J. A. A.; Gupta, H. N.; Bansal, B. B.; Srivastava, O. N. Int. J. Hydrogen Energy

2000, 25, 581-589.

7 Marinov, N. M.; Westbrook, C. K.; Pitz, W. J. “Transport Phenomena in Combustion.”

Chen, S. H., Ed. Eighth International Symposium on Transport Phenomena. 1996, p.

118.

8 Marinov, N. M.; Pitz, W. J.; Westbrook, C. K.; Castaldi, M. J. Senkan, S. Comb. Sci.

Tech. 196, 116-117, 211-287.

9 Tsang, W. J. Phys. Chem. Ref. Data, 1986, 15, 1087.

10 Sierens, R.; Rosseel, E. Proc. 12th World Hydrogen Conf., Buenos Aires, Argentina,

1998, 314-353.

11 Silva, E. P.; Gallo, W. L. R.; Szajner, J.; Amaral, E.G.; Bezerra, C. R. “State of the art in the use of hydrogen as an automotive fuel.” SAE Paper #931706E, 1993.

12 Cox, K. E.; Williamson, K. D. Hydrogen: Its Technology and Implications. Volumes

1-5. Boca Raton, FL: CRC Press, 1977.

13 Dulger, Z.; Ozcelik, K. R. Int. J. Hydrogen Energy 2000, 25, 895-897.

415

14 Tsolakis, A.; Megaritis, A.; Wyszynski, M. L. Energy Fuels 2003, 17, 1464-1473.

15 Tsolakis, A.; Megaritis, A. Int. J. Hydrogen Energy 2005, 30, 731-745.

16 Uykyr, C.; Henshaw, P. F.; Ting, D. S.-K.; Barron, R. M. Int. J. Hydrogen Energy

2001, 26, 265-273.

17 Hacochen, Y.; Shen, E. Int. J. Hydrogen Energy 1987, 12, 773-781.

18 Al-Janabi, H. A.-K. S.; Al-Janabi, M. A.- R. S. Int. J. Hydrogen Energy 1999, 24, 363-

375.

19 Sobiesiak, A.; Uykur, C.; Henshaw, P. F.; Ting, D. SAE Paper #2002-01-1710.

20 Henshaw, P.; D’Andrea, T.; Ting, D.; Sobiesiak, A. SAE Paper #2003-32-0011.

21 Naber, J. D.; Siebers, D. L. CSS/WSS/MNS of the Combustion Institute. 95S093,

1995.

22 Kumar, M. S.; Ramesh, A.; Nagalingam, B. Int. J. Hydrogen Energy 2003, 28, 1143-

1154.

23 Smith, G. P.; Golden, D. M.; Frenklach, M.; Moriarty, N. W.; Eiteneer, B.;

Goldenberg, M.; Bowman, C. T.; Hanson, R. K.; Song, S.; Gardiner, W. C., Jr.;

Lissianski, V. V.; Qin, Z. http://www.me.berkeley.edu/gri_mech/.

24 Curran, H.J.; Gaffuri, P.; Pitz, W.J.; Westbrook, C.K. Combust. Flame 1998, 114, 149-

177.

25 Curran, H.J.; Gaffuri, P.; Pitz, W.J.; Westbrook, C.K. Combust. Flame 2002, 129, 253-

280.

26 (a) Curran, H. J.; Pitz, W. J.; Westbrook, C. K.; Callahan, C. V.; Dryer, F. L. Proc.

Combust. Inst. 1998, 27, 379-387; (b) Huang, Y.; Sung, C. J.; Eng, J. A. Combust. Flame

416

2004, 139, 239-251; (c) Huang, Y.; Sung, C. J.; Eng, J. A. Proc. Combust. Inst. 2002, 29,

759-766.

27 Fournet, R.; Battin-Leclerc, F.; Glaude, P. A.; Judenherc, B.; Warth, V.; Come, G. M.;

Scacchi, G.; Ristori, A.; Pengloan, G.; Dagaut, P.; Cathonnet, M. Int. J. Chem. Kinet.

2001, 33, 574-586.

28 Kee, R. J.; Rupley, F. M.; Meeks, L. M.; Miller, J. A. A Fortran Chemical Kinetics

Package for the Analysis of Gas-Phase Chemical Kinetics; SAND96-8216; Sandia

National Laboratories: Livermore, CA, 1996. CHEMKIN Collection III, Reaction Design,

Inc.: San Diego, CA, 1999.

29 Reaction Design, http://www.reactiondesign.com

30 Grcar, J. F. “The TWOPNT Program for Boundary Value Programs.” Technical

Report SAND91-8230, Sandia National Laboratories, 1992.

31 Grcar, J. F.; Kee, R. J.; Smooke, M. D.; Miller, J. A. Proc. Combust. Inst. 1986, 21,

1773-1782.

32 Huang, Y.; Sung, C. J.; Eng, J. A. Combust. Flame 2004, 139, 239-251.

33 As the NOx submechanism is not part of either Curran mechanism, no NOx species were detected by CHEMKIN as possible emissions. Variance in NOx emissions is one of

the effects often cited with H2 augmentation; further calculations incorporating this set of

reactions would be instructive.

34 Mandilas, C.; Ormsby, M. P.; Sheppard, C. G.; Woolley, R. Proc. Combust. Inst. 2007,

31, 1443-1450.

417

BIBLIOGRAPHY

Aceves, S. M.; Flowers, D. L.; Martinez-Frias, J.; Smith, J. R.; Chen, J. Y. SAE Paper # 149614 (2002).

Aceves, S. M.; Smith, J. R.; Westbrook, C. K.; Iptz, W. J. J. Engin. Gas Turb. Power 1999, 121, 569-574.

Aceves, S.M.; Flowers, D.L.; Martinez-Frias, J.; Smith, J.R. SAE Paper #01-2077 (2001).

Akiyama, I.; Ogawa, M.; Takase, K.; Takamuku, T.; Yamaguchi, T.; Ohtori, N. J. Sol. Chem. 2004, 33, 797-809.

Albagi, A.; Oja, H.; Dubois, L. Environ. Lett. 1974, 6, 241.

Alfassi, Z. B.; Khaikin, G. I.; Neta, P. J. Phys. Chem. 1995, 99, 4544-4548.

Al-Janabi, H. A.-K. S.; Al-Janabi, M. A.- R. S. Int. J. Hydrogen Energy 1999, 24, 363- 375.

Amir,-Ebrahimi, V.; Choplin, A.; Demaison, J.; Roussy, G. J. Mol. Spectros. 1981, 89, 42-52.

Andino, J. M.; Smith, J. N.; Flagan, R. C.; Goddard, III, W. A.; Seinfield, J. H. J. Phys. Chem. 1996, 100, 10967-10980.

Aplincourt, P.; Ruiz-Lopez, M. F. J. Am. Chem. Soc. 2000, 122, 8990-8997.

Asaba, T.; Fujii, N. Proc. 13th Int. Combust. Symp. 1971, 155-164.

Asaba, T.; Fujii, N. Symp. (Int.) Combust., [Proc.] 14th, 1971, 155.

Atkinson, D. B.; Hudgens, J. W. J. Phys. Chem. A 1997, 101, 3901-3909.

Atkinson, R. Atmos. Environ. A 1990, 24A, 1-41.

Atkinson, R. J. Phys. Chem. Ref. Data 1997, 26, 215.

Atkinson, R. J. Phys. Chem. Ref. Data, Monograph 1, 1989, 1.

418 Atkinson, R. J. Phys. Chem. Ref. Data, Monograph 2, 1994, 1.

Atkinson, R.; Arey, J. Chem. Rev. 2003, 103, 4605-4638.

Atkinson, R.; Pitts, J. N. Chem. Phys. Lett. 1979, 63, 485.

Augspurger, J. D.; Dykstra, C. E.; Zwier, T. S. J. Phys. Chem. 1992, 96, 7252-7257.

Avramenko, L. I.; Kolesnikova, R. V.; Savinova, G. I. Bull. Acad. Sci. U.S.S.R. Div. Chem. Sci. (Engl. Transl.) 1965, 24-29.

Axworthy, A. E.; Dayan, V. H.; Martin, G. B. Fuel 1987, 57, 29.

Axworthy, A. Fuel 1987, 57, 29-35.

Babushok, V. I.; Tsang, W. Chemosphere 2003, 51, 1023-1029.

Bacskay, G. B.; Martoprawiro, M.; Mackie, J. C. Chem. Phys. Lett. 1998, 290, 391-398.

Bade, S.; Karim, G. A.; Wierzba, I. “Examination of operational limits in gas-fueled spark ignition engines.” SAE Paper #2000-01-1944- SP-1549.

Bajaj, P. N.; Fontijn, A. Combust. Flame 1996, 105, 239.

Baker, R. R.; Baldwin, R. R.; Walker, R. W. J. Chem. Soc., Faraday Trans. 1 1975, 71, 756.

Baker, R. R.; Bishop, L. J. J. Anal. Appl. Pyrol. 2005, 74, 145-170.

Baldwin, R. R.; Bennett, J. P.; Walker, R. W. J. Chem. Soc., Faraday Trans. 1 1980, 76, 1075.

Baldwin, R. R.; Cleugh, C. J.; Walker, R. W. J. Chem. Soc., Faraday Trans. 1 1976, 72, 1715.

Baldwin, R. R.; Pickering, I. A.; Walker, R. W. J. Chem. Soc., Faraday Trans. 1 1980, 76, 2374.

Barckholtz, C. A.; Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 1999, 103, 8108-8117.

Barckholtz, C.; Barckholtz, T. A.; Hadad, C. M. J. Am. Chem. Soc. 1999, 121, 491-500.

Barckholtz, C.; Barckholtz, T. A.; Hadad, C. M. J. Phys. Chem. A 2001, 105, 140-152.

Barker, J. R.; Ortiz, N. F. Int. J. Chem. Kinet. 2000, 33, 246.

419 Barone, V.; Cossi, M.; Tomasi, J. J. Chem. Phys. 1997, 107, 3210-3221.

Barone, V.; Cossi, M.; Tomasi, J. J. Comput. Chem. 1998, 19, 404-417.

Barry, N. J.; Fletcher, I. W.; Whitehead, J. C. J. Phys. Chem. 1986, 90, 4911.

Batt, L. Int. J. Chem. Kinet. 1979, 11, 977-993.

Batt, L.; Robinson, G. N. Int. J. Chem. Kinet. 1979, 11, 1045.

Bauer, E.; Magat, M. J. Phys. Radium 1938, 9, 319-330.

Baulch, D. L.; Cobos, C. J.; Cox, R. A.; Frank, P.; Hayman, G.; Just, T.; Kerr, J. A.; Murrells, T.; Pilling, M. J.; Troe, J.; Walker, R. W.; Warnatz, J. J. Phys. Chem. Ref. Data 1994, 23, 847-1033.

Bauschlichter, C. W., Jr.; Langhoff, S. R. Mol. Phys. 1999, 96, 471-476.

Beak, P. Acc. Chem. Res. 1977, 10, 186-192.

Bean, G. P. Tetrahedron 2002, 58, 9941.

Beck, S. M.; Brus, L. E. J. Chem. Phys. 1982, 76, 4700.

Becke, A. D. J. Chem. Phys. 1993, 98, 1372-1377.

Becke, A. D. J. Chem. Phys. 1993, 98, 5648;

Becke, A. D. Phys. Rev. A. 1988, 38, 3098-3100.

Becke, A.D. J. Chem. Phys. 1993, 98, 5648-5652.

Bent, D. V.; Haydon, E. J. Am. Chem. Soc. 1975, 97, 2599.

Bera, P. P.; Schaefer, H. F., III. Proc. Natl. Acad. Sci. USA. 2005, 102, 6698-6703.

Biggs, P.; Conosa-Mas, C. E.; Frachebod, J. M.; Shallcross, D. E.; Wayne, R. P. J. Chem. Soc., Faraday Trans. 1994, 90, 1205-1210.

Bittker, D. A. Prog. Energy Combust. Sci. 1991, 79, 49-72.

Bittner, J. D.; Howard, J. B. Proc. Combust. Inst. 1981, 18, 1105.

Bogdanov, B.; McMahon, T. B. J. Phys. Chem. A 2000, 104, 7871-7880.

Bowman, F. M.; Seinfeld, J. H. Prog. Energy Combust. Sci. 1995, 21, 387-417.

420 Bozzelli, J. W.; Dean, A. M. J. Phys. Chem. 1990, 94, 3313.

Bozzelli, J. W.; Pitz, W. J. Twenty-Fifth Symposium (International) on Combustion; The Combustion Institute: Pittsburgh, PA, 1994; 783-791.

Bozzelli, J. W.; Sheng, C. J. Phys. Chem. A 2002, 106, 1113-1121.

Braslavsky, S.; Heicklin, J. Chem. Rev. 1977, 77, 473-511.

Brezinsky, K.; Litzinger, T.; Glassman, I. Int. J. Chem. Kinet. 1984, 16, 1053-1061.

Bruinsma, O. S.; Tromp, P. J. J.; DeSauvage Nolting, H. J. J.; Mouljin, J. A. Fuel 1988, 67, 334-340.

Bunting, B.G. SAE Paper # 010872 (2006).

Burcat, A.; Kudchadker, S. Acta. Chim. Acad. Scient. Hung. 1979, 101, 249-258.

“Butanol.” Online. Internet. Available http://peswiki.com/index.php/Directory:Butanol. Accessed 15 July 2006.

“Butanol—Better than Ethanol as a Gasoline Substitute.” Online. Internet. Available http://alfin2300.blogspot.com/2006/06/butanol-better -than-ethanol-as.html. Accessed 30 June 2006.

Buth, R.; Hoyermann, K.; Seeba, J. 25th Symp. (Int’l) on Combustion; The Combustion Institute, 1994; 851-858.

Callam, C. S.; Singer, S. J.; Lowary, T. L.; Hadad, C. M. J. Am. Chem. Soc. 2001, 123, 11743-11754.

Calvert, J. G.; Atkinson, R.; Kerr, J. A.; Madronich, S.; Moortgat, G. K.; Wallington, T. J.; Yarwood, G. The Mechanisms of Atmospheric Oxidation of the Alkenes; Oxford University Press: New York, 2000.

Carey, F. A. Organic Chemistry, 5th Edition. New York: McGraw-Hill, 2003.

Carey, F. A.; Sundberg, R. J. Advanced Organic Chemistry: Part A, Structure and Mechanism. New York: Kluwer Academic/Plenum Publishers, 2000.

Carlsen, W. San Francisco Chronicle. 15 September 1997.

Carpenter, B. K. J. Am. Chem. Soc. 1993, 115, 9806-9807.

Carter, A. Classical and statistical thermodynamics. Upper Saddle River: Prentice Hall, 2001.

421 Carter, J. T.; Cook, D. B. J. Mol. Struct.: THEOCHEM 1991, 83, 111-122.

Casavecchia, P. Rep. Prog. Phys. 2000, 63, 355.

Cassanelli, P.; Cox, R. A.; Orlando, J. J.; Tyndall, G. S. J. Photochem. Photobiol. A 2006, 177, 109-115.

Cassanelli, P.; Johnson, D.; Cox, R. A. Phys. Chem. Chem. Phys. 2005, 7, 3702-3710.

Cavalli, F.; Geiger, H.; Barnes, I.; Becker, K.H. Environ. Sci. Technol. 2002, 36, 1263- 1270.

Chai, Y.; Pfefferle, L. D. Chem. Phys. Proc. Combust. 1995, 361-364.

Chai, Y.; Pfefferle, L. D. Fuel 1998, 77, 313-320.

Chan, C.-J.; Hamilton, I. P.; Pritchard, H. O. Faraday Trans. 1998, 94, 2303-2306.

Chandrasekhar, J.; Smith, S. F.; Jorgensen, W. L. J. Am. Chem. Soc. 1985, 107, 154-163.

Chang, H. M.; Jaffe, H. H. Chem. Phys. Lett. 1973, 23, 146.

Chang, H. M.; Jaffe, H. H.; Masmandis, C. A. J. Phys. Chem. 1975, 79, 1118.

Chase, R. “DuPont, BP join to make butanol.” USA TODAY. 30 June 2006.

Chen, C.-C.; Bozzelli, J. W. Fourth Joint Meeting of the U.S. Sections of the Combustion Institute: Western States, Central States, Eastern States. Philadelphia, PA. 20-23 March 2005.

Chen, C.-J.; Bozzelli, J. W. J. Phys. Chem. A 2000, 104, 4997.

Chen, P.; Colson, S. D.; Chupka, W. A.; Berson, J. A. J. Phys. Chem. 1986, 90, 2319;

Chen, Y.L.; Chou, T.-C. Ind. Eng. Chem. Res. 1996, 35, 2172-2176.

Choi, H.; Bise, R. T.; Neumark, D. M. J. Phys. Chem. A. 2000, 104, 10112-10118.

Chow, J. M.; Miller, A. M.; Elrod, M. J. J. Phys. Chem. A 2003, 107, 3040-3047.

Ciccioli, P.; Brancaleoni, E.; Frattoni, M.; Cecinato, A.; Pinciarelli, L. Anal. Lett. 2001, 34, 937-955.

Cizek, J. J. Chem. Phys. 1966, 45, 4256.

Clark, W. D.; Price, S. J. Can. J. Chem. 1970, 48, 1059.

422 Clark, W. D.; Price, S. J. Can. J. Chem. 1970, 48, 1059-1064.

Clifford, E. P.; Farrell, J. T.; DeSain, J. D.; Taatjes, C. A. J. Phys. Chem. A 2000, 104, 11549.

“Coal-Fired Power Generation.” Online. Internet. Available http://www.rst2.edu/ties/acidrain/IEcoal/how.htm Accessed 31 May 2006.

Colussi, A. J.; Singleton, D. L.; Irwin, R. S.; Cvetanovic, R. J. J. Phys. Chem. 1975, 79, 1900.

Colussi, A. J.; Zabel, F.; Benson, S. W. Int. J. Chem. Kin. 1977, 9, 161.

Conn, M. M.; Rebek, J., Jr. Chem. Rev. 1997, 97, 1647-1668.

Cool, T.; Qi, F.; McIlroy, A.; Westmoreland, P.R.; Law, M.; Poisson, L.; Peterka, D.; Ahmed, M. Rev. Sci. Instrum. 2005, 76, 941021-941027.

Cossi, M.; Barone, V. J. Chem. Phys. 1998, 109, 6246-6254

Cossi, M.; Barone, V.; Cammi, R.; Tomasi, J. Chem. Phys. Lett. 1996, 255, 327-335.

Cox, K. E.; Williamson, K. D. Hydrogen: Its Technology and Implications. Volumes 1-5. Boca Raton, FL: CRC Press, 1977.

Cox, R. A.; Tyndall, G. S. Chem. Phys. Lett. 1979, 65, 357-360.

Cramer, C.; Truhlar, D. Chem. Rev. , 1999, 99, 2161-2200.

Crosley, D. R. Laser Probes for Combustion Chemistry, American Chemical Society, ACS Symposium Series, 134, Washington, DC, 1980.

Cullis, C. F.; Norris, A. C. Carbon 1972, 10, 525-537.

Curran, H. J.; Gaffuri, P.; Pitz, W. J.; Westbrook, C. K. Combust. Flame 1998, 114, 149- 177.

Curran, H. J.; Gaffuri, P.; Pitz, W. J.; Westbrook, C. K. Combust. Flame 2002, 129, 253- 280.

Curran, H. J.; Jayaweera, T. M.; Pitz, W. J.; Westbrook, C. K. Western States Section of the Combustion Institute. Spring Meeting. Davis, CA, United States. 29 March 2004.

Curran, H. J.; Pitz, W. J.; Westbrook, C. K.; Callahan, C. V.; Dryer, F. L. Proc. Combust. Inst. 1998, 27, 379-387.

423 Curran, H.J.; Gaffuri, P.; Pitz, W.J.; Westbrook, C.K. Combust. Flame 1998, 114, 149- 177.

Curran, H.J.; Gaffuri, P.; Pitz, W.J.; Westbrook, C.K. Combust. Flame 2002, 129, 253- 280.

Cvetanovic, R. J. J. Phys. Chem. Ref. Data 1987, 16, 261.

Dagaut, P.; Pengloan, G.; Ristori, A. Phys. Chem. Chem. Phys. 2002, 4, 1846-1854.

Daimay, L.-V.; Colthup, N. B.; Fateley, W. G.; Grasselli, J. G. The Handbook of Infrared and Raman Characteristic Frequencies of Organic Molecules. Academic Press: Boston, 1991.

D'Anna, A.; Violi, A. Proc. Combust. Inst. 1998, 27, 425-433.

Das, L. M. Int. J. Hydrogen Energy 1990, 15, 833.

Davis, D. D.; Bollinger, W.; Fischer, S. J. Phys. Chem. 1975, 79, 293-294.

Dearden, D. V.; Beauchamp, J. L. J. Phys. Chem. 1985, 89, 5359.

Dechaux, J. C.; Delfosse, L. Combust. Flame 1979, 34, 169.

Demirba, A. Energ. Convers. Manage. 2002, 43, 2349-2356.

DeSain, J. D.; Clifford, E. P.; Taatjes, C. A. J. Phys. Chem. A 2001, 105, 3205-3213.

DeSain, J. D.; Klippenstein, S. J.; Miller, J. A.; Taatjes, C. A. J. Phys. Chem. A 2003, 107, 4415-4427.

DeSain, J. D.; Klippenstein, S. J.; Miller, J. A.; Taatjes, C. A. J. Phys. Chem. A 2004, 108, 7127-7128.

DeSain, J. D.; Taatjes, C. A.; Miller, J. A.; Klippenstein, S. J.; Hahn, D. K. Faraday Discuss. 2001, 119, 101-120.

DeSain, J.; Clifford, E.; Taatjes, C. J. Phys. Chem. A 2001, 105, 3205-3213.

DeSain, J.; Klippenstein, S.; Miller, J.; Taatjes, C. J. Phys. Chem. A 2001, 105, 4415- 4425.

Desiraju, G. R. Nature 2001, 412, 397-400.

Di Laudo, M.; Whittleton, S. R.; Wetmore, S. D. J. Phys. Chem. A 2003, 107, 10406- 10413.

424 Dimitrova, Y. J. Mol. Struct. (THEOCHEM) 1997, 391, 251-257.

DiPardo, J. “Outlook for Biomass Ethanol Production and Demand.” EIA Forecasts. Online. Internet. Available http://www.eia.doe.gov/oiaf/analysispaper/biomass.html. Accessed 6 July 2006.

Dobis, O.; Benson, S. W. J. Am. Chem. Soc. 1993, 115, 8798.

Doughty, A.; Mackie, J. C. J. Chem. Soc., Faraday Trans. 1994, 90, 541.

Doughty, A.; Mackie, J. C. J. Phys. Chem. 1992, 96, 10339-10348.

Doughty, A.; Mackie, J. C.; Palmer, J. M. 25th International Symposium on Combustion; The Combustion Institute: Pittsburgh, 1994; p. 893.

Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10339-10348.

Dulger, Z.; Ozcelik, K. R. Int. J. Hydrogen Energy 2000, 25, 895-897.

Ebert, K. H.; Ederer, H. J.; Schmidt, P. S., Baulch, D. L.; Cobos, C. J.; Cox, R. A.; Esser, C.; Frank, P.; Just, T.; Kerr, J. A.; Pilling, M. J.; Troe, J.; Walker, R. W.; Warnatz, J. J. Phys. Chem. Ref. Data 1992, 21, 411-429.

Ebrecht, J.; Hack, W.; Wagner, H. Ber. Bunsenges. Phys. Chem. 1989, 93, 619-623.

Eckbreth, A. C. Laser Diagnostics for Combustion Temperatures and Species, 2nd Ed., Gordon and Breach Publishers, Amsterdam B. V., Netherlands, 1996.

Eisele, F. L. Int. J. Mass Spectrom. Ion Phys. 1983, 54, 119-126.

El Bakali, A.; Dupont, L.; Lefort, B.; Lamoureux, N.; Pauwels, J. F.; Montero, M. J. Phys. Chem. A 2007, 111, 3907-3921.

El Dib, G.; Chakir, C; Roth, E.; Brion, J.; Daumont, D. J. Phys. Chem. A 2006, 110, 7848-7857.

Eliel, E.; Wilen, S. H.; Doyle, M. P. Basic Organic Stereochemistry. New York : Wiley- Interscience, 2001.

Ellison, G. B.; Davico, G. E.; Bierbaum, V. M.; DePuy, C. H. Int. J. Mass Spectrom. Ion Proc. 1996, 156, 109-131.

Ellison, G. B.; Davico, G. E.; Bierbaum, V. M.; DePuy, C. H. Int. J. Mass Spectrom. Ion Proc. 1996, 156, 109-131;

425 Elmaimouni, L.; Minetti, R.; Sawerysyn, J. P.; Devolder, P. Int. J. Chem. Kinet. 1993, 25, 399-413.

El-Nahas, A. M.; Navarro, M. V.; Simmie, J. M.; Bozzelli, J. W.; Curran, H. J.; Dooley, S.; Metcalfe, W. J. Phys. Chem. A. 2007, 111, 3727-3739.

Emdee, J. L.; Brezinsky, K.; Glassman, I. J. Phys. Chem. 1992, 96, 2151-2161.

Enami, S.; Yamanaka, T.; Nakayama, T.; Hashimoto, S.; Kawasaki, M.; Shallcross, D. E.; Nakano, Y.; Ishiwata, T. J. Phys. Chem. A 2007, 111, 3342-3348.

Endrenyi, L.; Le Roy, D. J. J. Phys. Chem. 1966, 70, 4081.

Energy Policy Act of 1992 (42 USC 10141 n). Online. Internet. Available http://thomas.loc.gov/cgi-bin/query/z?c102:H.R.776.ENR. Accessed 15 February 2007.

“Environmental Energy: Butanol Replaces Gasoline.” Online. Internet. Available http://www.butanol.com/. Accessed 30 June 2006.

Ergut, A.; Granata, S.; Jordan, J.; Carlson, J.; Howard, J. B.; Richter, H.; Levendis, Y. A. Combust. Flame 2006, 144, 757-772.

Erren, R. A.; Campbell, W. H. J. Inst. Fuel 1933, 6, 277-290.

Fadden, M. J.; Barckholtz, C.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 3004-3011.

Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6088-6094.

Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 6324-6331.

Fadden, M. J.; Hadad, C. M. J. Phys. Chem. A 2000, 104, 8121-8130.

Faulhaber, A. E.; Szpunar, D. E.; Kautzman, K. E.; Neumark, D. M. J. Phys. Chem. A. 2005, 109, 10239-10248.

Fenter, F.F.; Noziere, B.; Caralp, F.; Lesclaux, R. Int. J . Chem. Kinet. 1994, 26, 171-189.

Ferenac, M. A.; Davis, A. J.; Holloway, A. S.; Dibble, T. S. J. Phys. Chem. A 2003, 107, 63-72.

Finlayson-Pitts, B. J.; Pitts, Jr., J. N. Chemistry of the Upper and Lower Atmosphere: Theory, Experiments, and Applications. San Diego, CA : Academic Press, 2000.

Fish, A. Angew. Chem., Int. Ed. Engl. 1968, 7, 45.

426 Fisher, E.M.; Pitz, W.J.; Curran, H.J.; Westbrook, C.K. Proc. Combust. Inst. 2000, 28, 1579-1586.

Forst, W. Theory of Unimolecular Reactions; Academic Press: New York, 1973.

Fournet, R.; Battin-Leclerc, F.; Glaude, P. A.; Judenherc, B.; Warth, V.; Come, G. M.; Scacchi, G.; Ristori, A.; Pengloan, G.; Dagaut, P.; Cathonnet, M. Int. J. Chem. Kinet. 2001, 33, 574-586.

Frank, P.; Herzler, J.; Just, Th.; Wahl, C. Proc. 25th Intl. Symp. Combust. The Combustion Institute: Pittsburgh, PA, 1994; p 833.

Freedman, B.; Pryde, E. H. Proc. Int’l. Conf. on Plant and Vegetable Oils as Fuels, 1982, 117–22.

Frenklach, M. Phys. Chem. Chem. Phys. 2002, 4, 2028-2037.

Frenklach, M.; Wang, H. Proc. Combust. Inst. 1991, 23, 1559-1556.

Frerichs, H.; Schliephater, V.; Tappe, M.; Wagner, H. G. Z. Phys. Chem. 1990, 166, 157- 165.

Friderichsen, A. V.; Shin, E. J.; Evans, R. J.; Nimlos, M. R.; Dayton, D. C.; Ellison, G. B. Fuel 2001, 80, 1747-1755.

Fulle, D.; Dib, A.; Kiefer, J. H.; Zhang, Q.; Yao, J.; Kern, R. D. J. Phys. Chem. A 1998, 102, 7480-7486.

Fuls, J.; Hugo, F. J. C. Third International Conference On Energy Use Management, 1981. 1595–602.

Gail, S.; Thomson, M. J.; Sarathy, S. M.; Syed, S.A.; Dagaut, P.; Dievart, P.; Marchese, A. J.; Dryer, F. L. Proc. Combust. Inst. 2007, 31, in press.

Garrett, A. W.; Severance, D. L.; Zwier, T. S. J. Chem. Phys. 1992, 96, 7245-7258.

Garrett, A. W.; Zwier, T. S. J. Chem. Phys. 1992, 96, 7259-7267.

Gaussian 03, Revision C.02, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; ada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, . Y.; Morokuma, K.; Voth, G. A.; Salvador, P.;

427 Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; omaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A. Gaussian, Inc., Wallingford CT, 2004.

Gaussian 98, Revision A.11.03, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A., Jr.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; ammi, R.; Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.; Cui, Q.; Morokuma, K.; Salvador, P.; Dannenberg, J. J.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Baboul, A. G.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; omaromi, I.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Andres, J. L.; Gonzalez, C.; Head- Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian, Inc., Pittsburgh PA, 2001.

Ghigo, G.; Maranzana, A.; Tonachini, G. J. Chem. Phys. 2003, 118, 10575-10583.

Gilbert, R. G.; Smith, S. C. Theory of Unimolecular and Recombination Reactions; Blackwell Scientific Publications: Oxford, 1990.

Glassman, I., Ed. Combustion, 3rd Edition. Academic Press, San Diego, 1996.

Glaude, P. A.; Pitz, W. J.; Thomson, M. J. Proc. Combust. Inst. 2005, 30, 1111-1118.

Gomez, N.; Henon, E.; Bohr, F.; Devolder, P. J. Phys. Chem. A 2001, 105, 11204-11211.

Gong, X.; Xiao, H. J. Mol. Struct. (THEOCHEM) 1998, 453, 141-147.

Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1989, 90, 2154.

Gonzalez, C.; Schlegel, H. B. J. Phys. Chem. 1990, 94, 5523.

Good, D. A.; Francisco, J. S. J. Phys. Chem. A 2000, 104, 1171-1185.

Graboski, M. S.; McCormick, R. L. Prog. Energy Combust. Sci. 1998, 24, 125-164.

Graener, H.; Ye, T. Q.; Laubereau, A. J. Chem. Phys. 1989, 90, 3413-3416.

Grcar, J. F. “The TWOPNT Program for Boundary Value Programs.” Technical Report SAND91-8230, Sandia National Laboratories, 1992.

428 Grcar, J. F.; Kee, R. J.; Smooke, M. D.; Miller, J. A. Proc. Combust. Inst. 1986, 21, 1773-1782.

Grela, M. A.; Amorebieta, V. T.; Colussi, A. J. J. Phys. Chem. 1985, 89, 38-41.

Griffiths, J.; Halford-Maw, P.; Rose, D. Combust. Flame 1993, 95, 291-306. Roubaud, A.; Minetti, R.; Sochet, L. Combust. Flame 2000, 121, 535-541.

Gronert, S. J. Org. Chem. 2006, 71, 1209-1219.

Grotheer, H.-H.; Louw, R. Combust. Sci. Technol. 1998, 134, 45-63.

Grotheer, H.-H.; Louw, R. Symp. (Int.) Combust., [Proc.] 26th, 1996, 2405-2411.

Grovenstein, E., Jr.; Mosher, A. J. J. Am. Chem. Soc. 1970, 92, 3810-3812.

Gruenloh, C. J.; Florio, G. M.; Carney, J. R.; Hagemeister, F. C.; Zwier, T. S. J. Phys. Chem. A 1999, 103, 496-502.

Hacochen, Y.; Shen, E. Int. J. Hydrogen Energy 1987, 12, 773-781.

Harrison, A. G.; Honnen, L. R.; Dauben, H. J.; Lossing, F. P. J. Am. Chem. Soc. 1960, 82, 5593.

Hasson, A. S.; Tyndall, G. S.; Orlando, J. J. J. Phys. Chem. A 2000, 108, 5979-5989.

Hausmann, M.; Hebgen, P.; Homann, K. H. Symp. (Int.) Combust., [Proc.] 24th, 1992, 793-801.

Hayes, C. J.; Hadad, C. M. Combustion Pathways of the Alkyl-Substituted Heteroaromatics: Peroxy Radical Pathways of the Alkylated Azabenzenes, J. Phys. Chem. A, submitted.

Hayes, C. J.; Hadad, C. M. Combustion Pathways of the Alkyl-Substituted Heteroaromatics: Peroxy Radical Pathways of the Alkylated Five-Membered Heteroaromatics, J. Phys. Chem. A, submitted.

Hayes, C.J.; Hadad, C.M. Combustion Pathways of the Alkyl-Substituted Heteroaromatics: Bond Dissociation Enthalpies and Alkyl Group Fragmentations. J. Phys. Chem. A, submitted.

Hehre, W. J.; Radom, L.; Schleyer, P. v. R.; Pople, J. A. Ab Initio Molecular Orbital Theory. John Wiley & Sons: New York, 1986;

Hein, H.; Somnitz, H.; Hoffmann, A.; Zellner, R. Proceedings of the Second Workshop EUROTRAC-2 CMD, Karlsruhe, Germany, 1998.

429 Henshaw, P.; D’Andrea, T.; Ting, D.; Sobiesiak, A. SAE Paper #2003-32-0011.

Hickel, B. J. Phys. Chem. 1975, 79, 1054.

Hinchliffe, A.; Steinbank, R. E.; Ali, M. A. Theor. Chim. Acta 1966, 5, 95.

Hippler, H.; Troe, J. J. Phys. Chem. 1990, 94, 3803-3806.

Hodgson, D.; Zhang, H.-Y.; Nimlos, M. R.; McKinnon, J. T. J. Phys. Chem. A 2001, 105, 4316-4327.

Houser, T. J.; Hull, M.; Alway, R. M.; Biftu, T. Int. J. Chem. Kinet. 1980, 12, 569;

Houser, T. J.; McCarville, M. E.; Biftu, T. Int. J. Chem. Kinet. 1980, 12, 555;

Hoyermann, K.; Seeba, J. 25th Symp. (Int’l) on Combustion; The Combustion Institute, 1994; 851-858.

Huang, H.; Buekens, A. Organohalogen Compd. 1998, 36, 131-134.

Huang, H.; Fairweather, M.; Griffiths, J. F.; Tomlin, A. S.; Brad, R. B. Proc. Combust. Inst. 2005, 30, 1309-1316.

Huang, Y.; Sung, C. J.; Eng, J. A. Combust. Flame 2004, 139, 239-251.

Huang, Y.; Sung, C. J.; Eng, J. A. Proc. Combust. Inst. 2002, 29, 759-766.

Hughes, K. J.; Turányi,, T.; Clague, A.; Pilling, M. J. Int. J. Chem. Kinet., 2001, 33, 513- 538.

Hunziker, H. E.; Wendt, H. R. J. Chem. Phys. 1976, 64, 3488.

Ignatyev, I. S.; Xie, Y.; Allen, W. D.; Schaefer, H. F., III. J. Chem. Phys. 1997, 107, 141. Stark, M. S. J. Am. Chem. Soc. 2000, 122, 4162.

Ignatyev, I. S.; Xie, Y.; Allen, W. D.; Schaefer, H. F., III. J. Chem. Phys. 1997, 107, 141- 155.

Ikeda, E.; Mackie, J. J. Anal. Appl. Pyrolysis 1995, 34, 47-63.

Ilich, P.; Hemann, C. F.; Hille, R. J. Phys. Chem. B 1997, 101, 10923-10938.

IUPAC; http://www.iupac-kinetic.ch.cam.ac.uk

Jeffrey, G. A.; Saenger, W. Hydrogen Bonding in Biological Structures. Springer-Verlag: Berlin, 1991.

430 Jenkin, M. E.; Saunders, S. M.; Wagner, V.; Pilling, M. J. Atmos. Chem. 2003, 3, 181- 193.

Jitariu, L. C.; Jones, L. D.; Robertson, S. H.; Pilling, M. J.; Hillier, I. H. J. Phys. Chem. A 2003, 107, 8607-8617.

Johnson, B.G.; Gill, P.M.W.; Pople, J.A. J. Chem. Phys. 1993, 98, 5612-5626.

Johnson, C. R.; Ludwig, M.; Asher, S. A. J. Am. Chem. Soc. 1986, 108, 905.

Johnston, L. J.; Mathivanan, N.; Negri, F.; Siebrand, W.; Zerbetto, F. Can. J. Chem. 1993, 71, 1655.

Jones, J.; Bacskay, G. B.; Mackie, J. C.; Doughty, A. J. Chem. Soc., Faraday Trans. 1995, 91, 1587.

Jungkamp, T. P. W.; Smith, J. N.; Seinfeld, J. H. J. Phys. Chem. A 1997, 101, 4392-4401.

Just, G. M. P.; Sharp, E. N.; Zalyubovsky, S. J.; Miller, T. A. Chem. Phys. Lett. 2006, 417, 378-382.

Kaiser, E. W. J. Phys. Chem. 1995, 99, 707.

Kaiser, E. W. J. Phys. Chem. A 1998, 102, 5903-5906.

Kaiser, E. W. J. Phys. Chem. A 2002, 106, 1256.

Kaiser, E. W.; Lorkovic, I. M.; Wallington, T. J. J. Phys. Chem. 1990, 94, 3352.

Kaiser, E. W.; Rimai, L.; Wallington, T. J. J. Phys. Chem. 1989, 93, 4094.

Kaiser, E. W.; Wallington, T. J.; Andino, J. M. Chem. Phys. Lett. 1990, 168, 309.

Kan, C. S.; Calvert, J. G. Chem. Phys. Lett. 1979, 63, 111-114.

Kanno, H.; Honshoh, M.; Yoshimura, Y. J. Solution Chem. 2000, 29, 1007-1016.

Karim, G. A. Int. J. Hydrogen Energy 2003, 28, 569-577.

Kassel, L. S. J. Phys. Chem. 1928, 32, 1065-1079.

Kee, R. et al. Chemically Reacting Flow: Theory and Practice. John Wiley and Sons: Hoboken, NJ, 2003.

Kee, R. J.; Rupley, F. M.; Meeks, L. M.; Miller, J. A. A Fortran Chemical Kinetics Package for the Analysis of Gas-Phase Chemical Kinetics; SAND96-8216; Sandia

431 National Laboratories: Livermore, CA, 1996. CHEMKIN Collection III, Reaction Design, Inc.: San Diego, CA, 1999.

Kern, R. D.; Yong, J. N.; Kiefer, J. H.; Shah, J. N. 16th Int. Symp. Shock Tubes Waves 1987, 43;

Kiefer, J. H.; Mizerka, L. J.; Patel, M. R.; Wei, H. C. J. Phys. Chem. 1985, 89, 2013- 2019.

Kiefer, J.; Zhang, Q.; Kern, R.; Yao, J.; Jursic, B. J. Phys. Chem. A 1997, 101, 7061- 7073.

Kiefer, J.; Zhang, Q.; Kern, R.; Yao, J.; Jursic, B. J. Phys. Chem. B 1997, 101, 7061.

Kikuchi, O.; Hondo, Y.; Morihashi, K.; Nakayama, M. Bull. Chem. Soc. Jpn. 1988, 61, 291.

Kim, D. H.; Mulholland, J. A. Environ. Sci. Technol. 2005, 39, 5831-5836.

Kim, D. H.; Mulholland, J. A.; Ryu, J.-Y. Organohalogen Compd. 2004, 66, 1026-1032.

Kim, D. H.; Mulholland, J. A.; Ryu, J.-Y. Proc. Combust. Inst. 2005, 30, 1245-1253.

Kim, P.; Lee, J. H.; Bonanno, R. J.; Timmons, R. B. J. Chem. Phys. 1973, 59, 4593-4601.

King, R. O.; Rand, M. Canad. J. Technol. 1955, 33, 445-469.

Klein, E.; Lukes, V.; Cibulkova, Z.; Polovkova, J. J. Mol. Struct. 2006, 758, 149-159.

Klotz, S. D.; Brezinsky, K.; Glassman, I. Proceedings of the Combustion Institute; 1998, 27. 337-344.

Knox, J. H. Combust. Flame 1965, 9, 297.

Knox, J. H.; Wells, C. H. J. Trans. Faraday Soc. 1963, 59, 2786.

Knox, J. H.; Wells, C. H. J. Trans. Faraday Soc. 1963, 59, 2801.

Knutti, R.; Buhler, R. E. Chem. Phys. 1975, 7, 229-243.

Ko, T.; Adusei, G. Y.; Fontijn, A. J. Phys. Chem. 1991, 95, 8745.

Kohn, D. W.; Clauberg, H.; Chen, P. Rev. Sci. Instrum. 1992, 63, 4003.

Konnov, A.A., 28th Symposium (Int.) on Combustion, Edinburgh, 2000. Abstr. Symp. Pap. p. 317.

432 Krauss, M.; Osman, R. J. Phys. Chem. 1995, 99, 11387-11391.

Kroner, S. M.; DeMatteo, M. P.; Hadad, C. M.; Carpenter, B. K. J. Am. Chem. Soc. 2005, 127, 7466-7473.

Kulkarni, M. G.; Dalai, A. K. Ind. Eng. Chem. Res. 2006, 45, 2901-2913.

Kumar, M. S.; Ramesh, A.; Nagalingam, B. Int. J. Hydrogen Energy 2003, 28, 1143- 1154.

Labanowski, J. W.; Andzelm, J. Density Functional Methods in Chemistry; Springer: New York, 1991.

Laidler, K. J. Chemical Kinetics; Harper & Row, Publishers, Inc.: New York, 1987

Lalanne, P.; Andanson, J. M.; Soetens, J.-C.; Tassaing, T.; Danten, Y.; Besnard, M. J. Phys. Chem. A 2004, 108, 3902-3909.

Land, E. J.; Porter, G.; Strachan. Trans. Faraday Soc. 1961, 57, 1885.

Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B 1998, 37, 785-789.

Leidreiter, H. I.; Wagner, H. G. Z. Phys. Chem. 1989, 165, 1.

Leidreiter, H. I.; Wagner, H. Gg. Z. Z. Phys. Chem. Neue Floge 1987, 153, 99;

Lemieux, P. M.; Ryan, J. V. Air and Waste 1993, 43, 1106-1115.

Lendvay, G.; Viskolcz, B.; J. Phys. Chem. A 1998, 102, 10777-10786.

Lesar, A.; Hodoscek, M.; Drougas, E.; Kosmas, A. M. J. Phys. Chem. A 2006, 110, 7898- 7903.

Lesclaux, R. NATO Science Series, IV: Earth and Environmental Sciences 2002, 16, 79- 84.

Lesclaux, R.; Lightfoot, P.D.; Veyret, B. Poll. Atmos. 1991, 10, 69-79.

Levendis, Y. A.; Atal, A.; Carlson, J. B. Combust. Sci. Technol. 1998, 134, 407-431.

Lifshitz, A.; Tamburu, C.; Suslensky, A. J. Phys. Chem. 1989, 93, 5802-5808.

Lightfoot, P. D. Atmos. Environ. 1992, 26A, 180

Lightfoot, P. D.; Lesclaux, R.; Veyret, B. J. Phys. Chem. 1990, 94, 700-707.

433 Lin, C. Y.; Lin, M. C. J. Phys. Chem. 1986, 90, 425-431.

Lin, C.-Y.; Lin, H.-A.; Hung, L.-B. Fuel 2006, 85, 1743-1749.

Lin, C.-Y.; Lin, M. C. Int. J. Chem. Kinet. 1985, 17, 1025.

Lin, M. C.; Mebel, A. M. J. Phys. Org. Chem. 1995¸ 8¸407-420.

Linstedt, R. P. Combust. Sci. Technol., 1996, 120, 119-167.

Lipinsky, E. S.; McClure, T. A.; Wagner, C. K., Kresovich, S., Otis, J. L.; Applebaum, H. R. Proceedings of the International Conference on Plant and Vegetable Oils as Fuels, 1982. 1–10.

Liu, D.; Ma, G.; Levering, L. M.; Allen, H. C. J. Phys. Chem. B 2004, 108, 2252-2260.

Liu, J.; Hadad, C. M.; Platz, M. S. Org. Lett. 2005, 7, 549-552.

Liu, R.; Morokuma, K.; Mebel, A.M.; Lin, M.C. J. Phys. Chem. 1996, 100, 9314-9322.

Lossing, F. P. Ann. New York Acad. Sci. 1957, 67, 499-517.

Louw, R.; Grotheer, H.-H.; Kanters, M. J.; Wiater, I. Organohalogen Compd., 1996, 27, 26-29.

Lynch, B. J.; Fast, P. L.; Harris, M.; Truhlar, D. G. J. Phys. Chem. A 2000, 104, 4811.

Ma, F.; Hanna, M. A. Bioresour. Technol. 1999, 70, 1-15.

Ma, G.; Allen, H. C. J. Phys. Chem. B 2003, 107, 6343-6349.

Mackie, J. C.; Colket, M. B.; Nelson, P. F.; Esler, M. Int. J. Chem. Kinet. 1991, 23, 733- 760.

Mackie, J.; Colket, M.; Nelson, P. J. Phys. Chem. 1990, 94, 4099-4106.

Madronich, S.; Felder, W. J. Phys. Chem. 1985, 89, 3556-3561.

Mah, D. A.; Cabrera, J.; Nation, H.; Ramos, M.; Sharma, S.; Nickolaisen, S. L. J. Phys. Chem. A 2003, 107; 4354-4359.

Mammone, J. F.; Sharma, S. K. J. Phys. Chem. 1980, 84, 3130-3134.

Mandilas, C.; Ormsby, M. P.; Sheppard, C. G.; Woolley, R. Proc. Combust. Inst. 2007, 31, 1443-1450.

434 Marcus, R. A. J. Chem. Phys. 1952, 20, 359-364.

Maricq, M. M.; Szente, J. J. J. Phys. Chem. 1996, 100, 12374-12379.

Maricq, M. M.; Szente, J. J. J. Phys. Chem. A, 2000, 104, 7239-7245.

Marinov, N. M. Int. J. Chem. Kinet. 1999, 31, 183-220.

Marinov, N. M.; Castaldi, M. J.; Melius, C. F.; Tsang, W. Combust. Sci. Technol. 1997, 128, 295-342.

Marinov, N. M.; Pitz, W. J.; Westbrook, C. K.; Castaldi, M. J. Senkan, S. Comb. Sci. Tech. 1996, 116-117, 211-287.

Marinov, N. M.; Pitz, W. J.; Westbrook, C. K.; Castaldi, M. J.; Senkan, S. M.; Melius, C. F. Combust. Flame 1998, 114, 192-213.

Marinov, N. M.; Westbrook, C. K.; Pitz, W. J. “Transport Phenomena in Combustion.” Chen, S. H., Ed. Eighth International Symposium on Transport Phenomena. 1996, p. 118.

Martinez, E.; Albaladejo, J.; Jiminez, E.; Notario, A.; Diaz de Mera, Y. Chem. Phys. Lett. 2000, 329, 191-199.

Matus, M. H.; Nguyen, M. T.; Dixon, D. A. J. Phys. Chem. A. 2007, 111, 113-126.

McCreery, R. L. Raman Spectroscopy for Chemical Analysis. John Wiley & Sons: New York, 2000.

McDonald, J. D.; Zielinska, B.; Fujita, E. M.; Sagebiel, J. C.; Chow, J. C.; Watson, J. G. Env. Sci. Technol. 2000, 34, 2080-2091.

McEnally, C.S.; Pfefferle, L.D. Proc. Combust. Inst. 2006, 30, 1363-1370.

McEnally, C.S.; Pfefferle, L.D. Proc. Combust. Inst. 30:1363-1370 (2005).

McMillen, D. F.; Golden, D. M. Annu. Rev. Phys. Chem. 1982, 33, 493-532.

Mebel, A. M.; Lin, M. C. J. Am. Chem. Soc. 1994, 116, 9577-9584.

Mebel, A. M.; Lin, M. C.; Yu, T.; Morokuma, K. J. Phys. Chem. A 1997, 101, 3189- 3196.

Melius, C. F.; Colvin, M. E.; Marinov, N. M.; Pitz, W. J.; Senkin, S. M. Proc. Combust. Inst. 1996, 26, 685.

435 Meot-Ner, M. M.; Lias, S. G. Binding Energies Between Ions and Molecules, and the Thermochemistry of Cluster Ions. In NIST Chemistry WebBook, NIST Standard Reference Database Number 69. Linstrom, P. J., Mallard, W. G., Eds. National Institute of Standards and Technology: Gaithersburg, MD, March 2003.

Mereau, R.; Rayez, M.-T.; Caralp, F.; Rayez, J.-C. Phys. Chem. Chem. Phys. 2003, 5, 4828-4833.

Merle, J. K.; Hadad, C. M. J. Phys. Chem. A 2004, 108, 8419-8433.

Merle, J. K.; Hayes, C. J.; Zalyubovsky, S. J.; Glover, B. G.; Miller, T. A.; Hadad, C. M. J. Phys. Chem. A. 2005, 109, 3637-3646.

Metcalfe, W. K.; Dooley, S.; Curran, H. J.; Simmie, J. M.; El-Nahas, A. M.; Navarro, M. V.2007, 111, 4001-4014;

Meyers, R. A. Coal Structure; Academic Press, New York, 1982.

Miller, J. A.; Klippenstein, S. J. J. Phys. Chem. A, 2006, 110, 10528 -10544.

Miller, J. A.; Klippenstein, S. J.; Robertson, S. H. Proc. Combust. Inst. 2000, 28, 1479.

Miller, J. A.; Melius, C. F. Combust. Flame 1992, 91, 21-39.

Miller, J. A.; Pilling, M. J.; Troe, J. Proc. Combust. Inst. 2005, 30, 43-88.

Mirvish, S.S. Cancer Lett. 1995, 93, 17-48.

Molto, J.; Font, R.; Conesa, J.A. Energ. Fuels 2006, 20, 1951-1958.

Monks, P. S. Chem. Soc. Rev. 2005, 34, 376-395.

Muller-Markgraf, W.; Troe, J. J. Phys. Chem. 1988, 92, 4899 – 4905.

Muller-Markgraf, W.; Troe, J. J. Phys. Chem. 1988, 92, 4914-4922.

Muralha, V.; dos Santos, R.; Simoes, J. J. Phys. Chem. A 2004, 108, 936-942.

Naber, J. D.; Siebers, D. L. CSS/WSS/MNS of the Combustion Institute. 95S093, 1995.

Nag, P.; Song, K. H.; Litzinger, T. A.; Haworth, D. C. Int. J. Engine Res. 2001, 2, 163- 175.

Naik, C.; Carstensen, H.-H.; Dean, A. M. Proceedings of the Third Joint Meeting of the U.S. Sections of the Combustion Institute, Chicago, IL, 2003.

436 Najt, P. M.; Foster, D.E. SAE Paper #830264 (1983).

Nam, P. C.; Nguyen, M. T.; Chandra, A. J. Phys. Chem. A 2005, 109 10342 –10347.

Nguyen, T. L.; Peeters, J.; Vereecken, L. J. Phys. Chem. A. 2007, 111, 3836-3849.

Nicovich, J. M.; Ravishankara, A. R. J. Phys. Chem. 1984, 88, 2534-2541.

Noguchi, M.; Tanaka, Y.; Tanaka, T.; Takeuchi, Y. SAE Paper #790840 (1979).

Noziere, B.; Lesclaux, R.; Hurley, M. D.; Dearth, M. A.; Wallington, T. J. J. Phys. Chem. 1994, 98, 2864-2873.

Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A., Jr. J. Chem. Phys. 1996, 104, 2598-2619;

Olivella, S.; Sole, A.; Garcia-Raso, A. J. Phys. Chem. 1995, 99, 10549-10556.

Onishi, S.; Jo, S.; Shoda, K.; Jo, P. D.; Kato, S. SAE Paper #790501 (1979).

Onsager, L. J. Am. Chem. Soc. 1936, 58, 1486-1493.

Organ, P. P.; Mackie, J. C. Chem. Soc. Faraday Trans. 1991, 87, 815-823.

Orlando, J. J.; Tyndall, G. S.; Bilde, M.; Ferronato, C.; Wallington, T. J.; Vereecken, L.; Peeters, J. J. Phys. Chem. A 1998, 102, 8116;

Orozco, M.; Luque, F. J. Chem. Rev. 2000, 100, 4187-4226.

Ozawa, T.; Ishiguro, Y.; Toyoda, K.; Nishimura, M.; Sasahara, T.; Doi, T. Chem. Sensors 2004, 20, 56-57.

Pacansky, J.; Waltman, R. J.; Barnes, L. A. J. Phys. Chem. 1993, 97, 10694.

Page, M.; Lin, M. C.; He, Y.; Choudhury, T. K. J. Phys. Chem. 1989, 93, 4404-4408.

Pan, X. M.; Fu, Z.; Li, Z. S.; Sun, C. C.; Sun, H.; Su, Z. M.; Wang, R. S. Chem. Phys. Lett. 2005, 409, 98-104.

Pang, X.; Feng, W.; Wang, Y.; Zhang, S. Wuli Huaxue Xuebao 1996, 12, 391-395.

Park, J.; Dyakov, I. V.; Lin, M. C. J. Phys. Chem. A 1997, 101, 8839-8843.

Parr, R. G.; Yang, W. Density Functional Theory in Atoms and Molecules; Oxford University Press: New York, 1989.

437 Pate, C. T.; Sprung, J. L.; Pitts, Jr., J. N. Org. Mass. Spec. 2005, 11, 552-555.

Pavanello, S.; Simioli, P.; Lupi, S.; Gregorio, P.; Clinfero, E. Cancer Epidemiology, Biomarkers and Prevention 2002, 11, 998.

Perry, R. A.; Atkinson, R.; Pitts, J. N., Jr. J. Phys. Chem. 1977, 81, 296-304.

Petersson, G. A.; Frisch, M. J. J. Phys. Chem. A 2000, 104, 2183.

Pilling, M. J.; Robertson, S. H.; Seakins, P. W. J. Chem. Soc., Faraday Trans. 1995, 91, 4179.

Pitz, W. J.; Seiser, R.; Bozzelli, J. W.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin- Leclerc, F.; Seshadri, K.; Westbrook, C. K. 30th International Symposium on Combustion; July 2004.

Pitz, W. J.; Seiser, R.; Bozzelli, J.W.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin- Leclerc, F.; Seshadri, K.; Westbrook, C.K. 2nd Joint Meeting of the U.S. Section of the Combustion Institute: Oakland, CA. March 25-28, 2001.

Pitz, W.J.; Seiser, R.; Bozzelli, J.W.; Seshadri, K.; Chen, C. J.; Da Costa, I.; Fournet, R.; Billaud, F.; Battin-Leclerc, F.; Westbrook, C. K. Thirtieth Symposium (International) on Combustion. The Combustion Institute; Chicago, IL: July 25-30, 2004.

Pople, J. A.; Scott, A. P.; Wong, M. W.; Radom, L. Isr. J. Chem. 1993, 33, 345-350.

Pribble, R. N.; Hagemeister, F. C.; Zwier, T. S. J. Chem. Phys. 1997, 106.

Provencal, R. A.; Casaes, R. N.; Roth, K.; Paul, J. B.; Chapo, C. N.; Saykally, R. J.; Tschumper, G. S.; Schaefer III, H. F. J. Phys. Chem. A 2000, 104, 1423-1429.

Pullin, D.; Andrews, L. J. Mol. Struct. 1982, 95, 181.

Pushkarsky, M. B.; Zalyubovsky, S. J.; Miller, T. A. J. Chem. Phys. 2000, 112, 10695.

Qin, Z.; Lissianski, V. V.; Yang, H.; Gardiner, W. C.; Davis, S. G.; Wang, H. Proc. Combust. Inst. 2000, 28, 1663-1669.

Quelch, G. E.; Gallo, M. M.; Schaefer, H. F., III. J. Am. Chem. Soc. 1992, 114, 8239.

Raffaelli, A.; Saba, A. Mass. Spec. Rev. 2003, 22, 318-331.

Ramey, D. “Advances in Biofuels: Alternative Energy and New Markets for Farmers.” Online. Internet. Available http://www.lightparty.com/Energy/Butanol.html. Accessed 24 July 2006.

438 Ramey, D.; Yang, S.-T. Production of Butyric Acid and Butanol from Biomass. U.S. Department of Energy: Morgantown, WV, 2004. Online. Internet. Available http://www.osti.gov/bridge/servlets/purl/843183-rqlY2a/843183.PDF. Accessed 15 July 2006.

Ramirez, F. J.; Tunon, I.; Silla, E. J. Phys. Chem. B 1998, 102, 6290-6298.

Rao, V. S.; Skinner, G. B. J. Phys. Chem. 1984, 88, 5990-5995.

Rawlins, W. T.; Tanzawa, T.; Schertzer, S. P.; Kreck. R. H. Final Report to DOE, 1983, Contract No. DEAC228OPC30392;

Reaction Design, http://www.reactiondesign.com

Reed, A. E.; Weinstock, R. B.; Weinhold, F. J. Chem. Phys. 1985, 83, 735-746.

Reed, A. E; Curtiss, L. A.; Weinhold, F. Chem. Rev. 1988, 88, 899-926.

Rice, O. K.; Ramsperger, H. C. J. Am. Chem. Soc. 1928, 50, 617-620.

Richter, H.; Howard, J. B. Phys. Chem. Chem. Phys. 2002, 4, 2038-2055.

Richter, H.; Howard, J. B. Prog. Energy Combust. Sci. 2000, 26, 565.

Rienstra-Kiracofe, J. C.; Allen, W. D.; Scaefer, III, H. F. J. Phys. Chem. A 2000, 104, 9823-9840.

Rivail, J. L.; Rinaldi, D.; Dillet, V. Mol. Phys. 1996, 89, 1521-1529.

Robinson, P. J.; Holbrook, K. A. Unimolecular Reactions; Wiley: New York, 1972.

Roebber, J. L. J. Chem. Phys. 1962, 37, 1974.

Rohrs, H. W.; Wickham-Jones, T.; Ellison, G. B.; Berry, D.; Argrow, B. M. Rev. Sci. Instrum. 1995, 66, 2430-2441.

Ross, P. L.; Johnston, M. V. J. Phys. Chem. 1995, 99, 16507-16510.

Rupper, P.; Sharp, E. N.; Tarczay, G.; Miller, T. A. J. Phys. Chem. A 2007, 111, 832-840.

Ruscic, B.; Berkowitz, J. Preprints of Papers, Division of Fuel Chemistry. 1991, 36, 1571-1575. 202nd American Chemical Society National Meeting, New York City, August 25-30, 1991.

Sanders, W. A. Gov. Rep. Announce. Index (U.S.) 1986, 86(15), Abstr. No. 633-265.

439 Sarathy, S. M.; Gail, S.; Syed, S. A.; Thomson, M. J.; Dagaut, P. Proc. Combust. Inst. 2007, in press.

Sassi, P.; Morresi, A.; Paolantoni, M.; Cataliotti, R. S. J. Mol. Liq. 2002, 96-97, 363-377.

Sauer, M. C., Jr.; Mani, I. J. Phys. Chem. 1970, 74, 59-63.

Sauer, M. C., Jr.; Ward, B. J. Phys. Chem. 1967, 71, 3971-3983.

Saunders, S. M.; Jenkin, M. E.; Derwent, R. G.; Pilling, M. J. Atmos. Chem. 2003, 3, 161- 180.

Schlegel, H. B. J . Chem. Phys. 1986, 84, 4530-4534.

Schmoltner, A.-M.; Anex, D. S.; Lee, Y. T. J. Phys. Chem. 1992, 96, 1236.

Schoebel, A. Wissenschaftliche Berichte-Forschungszentrum Karlsruhe. 2000, i-vii, 1- 143.

Schuler, R. H.; Buzzard, G. K. Int. J. Radiat. Phys. Chem. 1976, 8, 563.

Schwab, A. W.; Bagby, M. O.; Freedman, B. Fuel 1987, 66, 1372-1378;

Schwab, A. W.; Dykstra, G. J.; Selke, E.; Sorenson, S. C.; Pryde, E. H. JOACS 1988, 65, 1781-1786.

Schwartz, W. R.; McEnally, C. S.; Pfefferle, L. D. J. Phys. Chem. A 2006, 110, 6643- 6648.

Scott, A. P.; Radom, L. J. Phys. Chem. 1996, 100, 16502-16513.

Scott, M.; Walker, R. W. Combust. Flame 2002, 129, 365-377.

Sebbar, N.; Bockhorn, H.; Bozzelli, J. W. J. Phys. Chem. A 2005, 109, 2233-2253;

Sebbar, N.; Bockhorn, H.; Bozzelli, J. W. Phys. Chem. Chem. Phys. 2002, 15, 3691- 3703;

Semenov, N. N. Chemical Kinetics and Chain Reactions. The Clarendon Press: Oxford, 1973.

Semenov, N. N. Some Problems in Chemical Kinetics and Reactivity. Princeton University Press: Princeton, NJ 1958;

Sendt, K.; Bacskay, G.B.; Mackie, J. C. J. Phys. Chem. 2000, 104, 1861-1875.

440 Sethuraman, S.; Senkan, S. M.; Gutman, D. Combust. Sci. Technol. 1992, 82, 13-30.

Sibener, S. J.; Buss, R. J.; Casavecchia, P.; Hirooka, T.; Lee, Y. T. J. Chem. Phys. 1980, 72, 4341.

Sidhu, S.; Graham, J.; Stribech, R. Chemosphere 2001, 42, 681-690.

Sierens, R.; Rosseel, E. Proc. 12th World Hydrogen Conf., Buenos Aires, Argentina, 1998, 314-353.

Silva, E. P.; Gallo, W. L. R.; Szajner, J.; Amaral, E.G.; Bezerra, C. R. “State of the art in the use of hydrogen as an automotive fuel.” SAE Paper #931706E, 1993.

Simonaitis, R.; Heicklen, J. J. Phys. Chem. 1974, 78, 2417-2421.

Sinha, A.; Thomson, M. J. Combust. Flame 2004, 136, 548-556.

Smith, G. P.; Golden, D. M.; Frenklach, M.; Moriarty, N. W.; Eiteneer, B.; Goldenberg, M.; Bowman, C. T.; Hanson, R. K.; Song, S.; Gardiner, W. C., Jr.; Lissianski, V. V.; Qin, Z. http://www.me.berkeley.edu/gri_mech/.

Smith, K. L.; Smoot, L. D.; Fletcher, T. H.; Pugmire, R. J. The Structure and Reaction Processes of Coal; Plenum Press: New York, 1994.

Smith, L., Jr.; Evans, S. J. Chem. Soc. Perkin Trans. 2 2000, 7, 1541-1551.

Sobiesiak, A.; Uykur, C.; Henshaw, P. F.; Ting, D. SAE Paper #2002-01-1710.

Somnitz, H.; Zellner, R. Z. Phys. Chem. 2006, 220, 1029-1048.

Song, S.; Golden, D. M.; Hanson, R. K.; Bowman, C. T. J. Phys. Chem. A 2002, 106, 6094.

Srivastava, A.; Prasad, R. Renew. Sust. Energ. Rev. 2000, 4, 111-133.

Stephens, P.J.; Devlin, F.J.; Frisch, M.J. J. Phys. Chem. 1994, 98, 11623-11627.

Streit, G. E.; Spall, D. E.; Hall, J. H. Environ. Sci. Technol. 1987, 21, 519-525.

Sun, Y.; Huang, X.-R.; Li, Z.-S.; Sun, C.-C. Gao. Xue. Hua. Xue. 2002, 23, 278-282.

Sussingham, R. “Can Coal Be Clean?” Chemistry. The American Chemical Society. Spring 2007.

Suzuki, S.; Green, P. G.; Bumgarner, R. E.; Dasgupta, S.; Goddard III, W. A.; Blake, G. A. Science 1992, 257, 942-945.

441 Taatjes, C. A.; Hansen, N.; McIlroy, A. Miller, J. A.; Senosiain, J. P.; Klippenstein, S. J.; Qi, F.; Sheng, L.; Zhang, Y.; Cool, T. A.; Wang, J.; Westmoreland, P. R.; Law, M. E.; Kasper, T.; Kohse-Höinghaus, K. Science, 2005, 308, 1887-1889.

Taatjes, C. A.; Hansen, N.; Miller, J. A.; Cool, T. A.; Wang, J.; Westmoreland, P. R.; Law, M. E.; Kasper, T.; Kohse-Hoinghaus, K. Proc. Combust. Inst. 2005, 30, 43-49.

Taberes, F.L.; Gonzalez-Urena, A. J. Phys. Chem. 1983, 87, 4933-4936.

Tappe, M.; Schliephake, V.; Wagner, H. G. Z. Phys. Chem. 1989, 162, 129.

Tarczay, G.; Zalyubovsky, S. J.; Miller, T.A. Chem. Phys. Lett. 2005, 406, 81-89.

Tardy, D. C. Int. J. Chem. Kinet. 1974, 6, 291.

Terentis, A.; Doughty, A.; Mackie, J. J. Phys. Chem. 1992, 96, 10334-10339.

Thring, R. H. Homogeneous-Charge Compression-Ignition (HCCI) Engines. Warrendale, PA: SAE, 1989

Tokmakov, I. V.; Kim, G.-S.; Kislov, V. V.; Mebel, A. M.; Lin, M. C. J. Phys. Chem. A 2005, 109, 6114-6127.

Tomasi, J. Chem. Rev., 1994, 94, 2027-2094.

Tonokura, K.; Norikane, Y.; Koshi, M.; Nakano, Y.; Nakamichi, S.; Goto, M.; Hashimoto, S.; Kawasaki, M.; Andersen, M. P.; Sulbaek, M. P.; Hurley, M. D.; Wallington, T. J. J. Phys. Chem. A 2002, 106, 5908-5917.

Tripathi, G. N. R. In Time Resolved Spectroscopy; Clark, R. J. H., Hester, R. E., Eds.; Wiley: New York, 1989; p 157.

Tripathi, G. N. R.; Schuler, R. H. Chem. Phys. Lett. 1983, 98, 594.

Tripathi, G. N. R.; Schuler, R. H. J. Chem. Phys. 1984, 81, 113.

Tripathi, G. N. R.; Schuler, R. H. J. Phys. Chem. 1988, 92, 5129.

Tsang, W. J. Phys. Chem. Ref. Data, 1986, 15, 1087.

Tsang, W.; Bedanov, V.; Zachariah, M. R. Ber. Bunsen-Ges. Phys. Chem. 1997, 101, 491.

Tsang, W.; Lifshitz, A. Ann. Rev. Phys. Chem. 1990, 41, 559.

442 Tsang, W.; Walker, J. A.; Manion, J. A. Symp. (Int.) Combust., [Proc] 1998, 27th (1), 135.

Tsolakis, A.; Megaritis, A. Int. J. Hydrogen Energy 2005, 30, 731-745.

Tsolakis, A.; Megaritis, A.; Wyszynski, M. L. Energy Fuels 2003, 17, 1464-1473.

Tully, F. P.; Ravishankara, A. R.; Thompson, R. L.; Nicovich, J. M.; Shah, R. C.; Kreutter, N. M.; Wine, P. H. J. Phys. Chem. 1981, 85, 2262-2269.

“University Bolsters Major Endeavor to Solve Global Energy Problems,” Ohio State University News in Engineering. Online. Internet. Available http://engineering.osu.edu/nie/article.php?e=783&s=3&a=2. Accessed 15 February 2007.

Uykyr, C.; Henshaw, P. F.; Ting, D. S.-K.; Barron, R. M. Int. J. Hydrogen Energy 2001, 26, 265-273.

Van der Laarse, J. D.; van Leuven, H. C. E. Anal. Chim. Acta 1966, 34, 370-371.

Veldhuizen, R.; de Leeuw, S. W. J. Chem. Phys. 1996, 105, 2828-2836.

Venkat, C.; Brezinsky, K.; Glassman, I. Proceedings of the Combustion Institute; 1982, 19, 143-152.

Venkat, C.; Brezinsky, K.; Glassman, I. Symp. (Int.) Combust., 19th 1982, 143-152.

Vereecken, L.; Peeters, J. J. Chem. Phys. 2003, 119, 5159-5170.

Vereecken, L.; Peeters, J. J. Phys. Chem. A 1999, 103, 1768;

Vereecken, L.; Peeters, J.; Orlando, J. J.; Tyndall, G. S.; Ferronato, C. J. Phys. Chem. A 1999, 103, 4693.

Vereecken, L.; Peters, J. J. Chem. Phys. 2003, 119, 5159-5170.

Vovelle, C.; Delfau, J. L. NATO Science Series, Series C: Mathematical and Physical Sciences 2000, 183-203.

Wagner, A. F.; Slagle, I. R.; Sarzynski, D.; Gutman, D. J. Phys. Chem. 1990, 94, 1853.

Wagner, A. F.; Slagle, I. R.; Sarzynski, D.; Gutman, D. J. Phys. Chem. 1990, 94, 1853;

Wahner, A.; Zetsch, C. J. Phys. Chem. 1983, 87, 4945-4951.

Walker, J.; Tsang, W. J. Phys. Chem. 1990, 94, 3324-3327.

443 Walker, R. W. In Major Research Topics in Combustion; Hussaini, M. Y., Kumar, A., Voigt, R. G., Eds.; Springer-Verlag: New York, 1992; p 277.

Walker, R. W.; Morley, C. Basic Chemistry of Combustion. In Low-Temperature Combustion and Autoignition; Willing, M. J., Ed.; Elsevier: Amsterdam; New York, 1997; pp 1-124.

Wallington, T. J.; Dagaut, R.; Kurylo, M. J. Chem. Rev. 1992, 92, 667-710;

Wallington, T. J.; Kaiser, E. W.; Farrell, J. T. Chem. Soc. Rev. 2006, 35, 335-347.

Wallington, T. J.; Maricq, M. M.; Ellermann, T.; Nielsen, O. J.; J. Phys. Chem. 1992, 96, 982-986.

Wallington, T. J.; Nielsen, O. J.; Sehested, J. “Reactions of Organic Peroxy Radicals in the Gas Phase,” in Peroxyl Radicals, Ed. Alfassi, Z. John Wiley and Sons: New York, 1997.

Walsh, R. Chem. Rev. 1969, 69, 103-124.

Wang, H.; Frenklach, M. J. Phys. Chem. 1994, 98, 11465-11489.

Wang, H.; Frenklach, M. J. Phys. Chem. 1994, 98, 11465-11489;

Wang, R.C.; Fang, T. SAE Paper #011519 (2006).

Wang, Z.; Wang, J.; Richter, H.; Howard, J. B.; Carlson, J.; Levendis, Y. A. Energy Fuels, 2003, 17, 999-1013.

Ward, B. Spectrochim. Acta A 1968, 24, 813.

Warnatz, J. “Rate coefficients in C/H/O system.” Ch. 5 in Combustion Chemistry, W. C. Gardiner, Jr., Ed., Springer-Verlag, New York, 1984.

Warnatz, J.; Maas, U.; Dibble, R. W. Combustion: physical and chemical fundamentals, modeling and simulation, experiments, pollutant formation. New York: Springer, 2001.

Watkins, K. W. J. Am. Chem. Soc. 1971, 93, 6355.

Watkins, K. W. J. Phys. Chem. 1973, 77, 2938.

We note that the conformer of Q(1,4s)OOH formed via this step different than that formed via direct 1,4-H atom transfer in the initial 1-peroxy propanol; hence, the energy of the two conformers differs by ~1.5 kcal/mol.

Weisman, J. L.; Head-Gordon, M. J. Am. Chem. Soc. 2001, 123, 11686-11694.

444 Wen, Z.; Li, Z.; Shang, Z.; Cheng, J.-P. J. Org. Chem. 2001, 66, 1466.

Westbrook, C. K. 28th Int’l Symp. Comb. 2000, 1563-1577.

Westbrook, C. K.; Mizobuchi, Y.; Poinsot, T. J.; Smith, P. J.; Warnatz, J. Proc. Combust. Inst. 2005, 30, 125-157.

Westbrook, C. K.; Pitz, W. J.; Curran, H. J. J. Phys. Chem. A 2006, 110, 6912-6922.

Westmoreland, P. R.; Dean, A. M.; Howard, J. B.; Longwell, J. P. J. Phys. Chem. 1989, 93, 8171.

Whitesides, G. M.; Mathias, J. P.; Seto, C. T. Science 1991, 254, 1312-1319.

WHO (1997). The World Health Report. World Health Organization, Geneva, Switzerland.

Wiater-Protas, I.; Louw, R. Eur. J. Org. Chem. 2001, 20, 3945-3952.

Wiberg, K. B.; Nakaji, D.; Breneman, C. M. J. Am. Chem. Soc. 1989, 111, 4178-4190.

Wigner, E. P. Z. Phys. Chem. Abst. B 1932, 19, 203.)

Williams, A.; Pourkashanian, M.; Jones, J.M.; Rowlands, L. J. Inst. Energy, 1997, 70, 102-113.

Williams, J. A.; Martin, F. L.;Muir, G. H.; Hewer, A.; Grover, P. L.; Phillips, D. H. Carcinogenesis 2000, 21, 1683-1689.

Wolfrom, J. Faraday Discuss. 2001, 119, 1-26.

Wollenhaupt, M.; Crowley, J. N. J. Phys. Chem. A 2000, 104, 6429-6438.

Woutersen, S.; Emmerichs, U.; Bakker, H. J. J. Chem. Phys. 1997, 107, 1483-1490.

Xie, H.; Wang, J.; Zhang, S.; Ding, Y.; Sun, C. J. Chem. Phys. 2006, 125, 124317- 124317-10.

Xing, J.-H., Nagai, Y., Kusuhara, M., and Miyoshi, A. J. Phys. Chem. A 2004, 108, 47, 10458- 10463.

Yamauchi, N.; Miyoshi, A.; Kosaka, K.; Koshi, M.; Matsui, H. J. Phys. Chem. A 1999, 103, 2723.

Yamin, J. A. A.; Gupta, H. N.; Bansal, B. B.; Srivastava, O. N. Int. J. Hydrogen Energy 2000, 25, 581-589.

445 Yang, K. J. Am. Chem. Soc. 1962, 84, 3795.

Yeung, L.Y.; Elrod, M. J. J. Phys. Chem. A 2003, 107, 4470-4477.

Yu, T.; Lin, M.C. J. Am. Chem. Soc. 1994, 116, 9571-9576.

Zalyubovsky, S. J.; Glover, B. G.; Miller, T. A.; Hayes, C. J.; Merle, J. K.; Hadad, C. M. J. Phys. Chem. A 2005, 109, 1308-1315.

Zalyubovsky, S. J.; Glover, B.; Miller, T. A.; Hayes, C. J.; Merle, J. K.; Hadad, C. M. J. Phys. Chem. A 2005, 107, 1308-1315.

Zalyubovsky, S.; Glover, B.; Miller, T.; Hayes, C.; Merle, J.; Hadad, C.; J. Phys. Chem. A, 2005, 107, 1308-1315.

Zaslonko, I. S.; Mozzhukhin, E. V.; Smirnov, V. N.; Tyurin, A. N. Kin.i Kat. 1991, 24, 131-153.

Zerefos, C.S.; Isaksen, I. S. A.; Ziomas, I. Chemistry and radiation changes in the ozone layer. Boston: Kluwer Academic Publishers, 2000.

Zhang, H.-Y.; McKinnon, J. T. Combust. Sci. Technol. 1995, 107, 261-300.

Zhang, X.; Maccarone, A. T.; Nimlos, M. R.; Kato, S.; Bierbaum, V. M.; Ellison, G. B.; Ruscic, B.; Simmonnett, A. C.; Allen, W. D.; Schaefer, III, H. F. J. Chem. Phys. 2007, 126, 44312.

Zheng, J.; Kwak, K.; Asbury, J.; Chen, X.; Piletic, I.R.; Fayer, M. D. Science 2005, 309, 1338-1343.

Zhenyi, C.; Xing, J.; Shuyuan, L.; Li, L. Energy Sources 2004, 26, 849-856.

Zhong, X.; Bozzelli, J. Int. J. Chem. Kinet. 1997, 29, 893-913.

Zhong, X.; Bozzelli, J. J. Phys. Chem. A 1998, 102, 3537-3555.

Zhou, D.; Nag, M.; Russel, A. L.; Read, D.; Rohrs, H. W.; Gross, M. L.; Gaspar, P. P. Sil. Chem. 2005, 3, 117-122.

Zwier, T. S. Annu. Rev. Phys. Chem. 1996, 47, 205-241.

446