PART I: PART II: 0—7639 LIBRARY Michigan State Unrvasby‘ This is to certify that the thesis entitled INTRAMOLECULAR CHARGE-TRANSFER QUENCHING OF EXCITED STATES BY SULFUR RADICAL B-CLEAVAGE VIA EXCITED STATESAND PHOTOGENERATED DIRADICALS presented by Michael Jeffrey Lindstrom ; has been accepted towards fulfillment of the requirements for Chemistry 4!” /£jor profeésor Ph. D. degree in PART I INTRAMOLECULAR CHARGE-TRANSFER QUENCHING OF EXCITED STATES BY SULFUR PART II RADICAL B-CLEAVAGE VIA EXCITED STATES AND PHOTOGENERATED DIRADICALS BY Michael Jeffrey Lindstrom A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1978 ABSTRACT PART I INTRAMOLECULAR CHARGE-TRANSFER QUENCHING OF EXCITED STATES BY SULFUR PART II RADICAL B-CLEAVAGE VIA EXCITED STATES AND PHOTOGENERATED DIRADICALS BY Michael Jeffrey Lindstrom PART I The photochemistry of various benzoylsulfides, sulfoxides, and sulfones was studied to determinelfluaeffect of chain length upon the rate of interaction of the excited benzoyl with the sulfur moiety. In general, the rate of charge-transfer quenching increases as the number of carbons between the donor and acceptor decreases, with the exception of the a-thioalkoxy- acetophenones, which quench more slowly than the B- and y-thioalkoxyketones. The observed trend is rationalized in terms of the ease of formation of various sized rings, which,jJ1turn, corresponds to approach of donor and accep- tor. Increasing the oxidation state of the sulfur group also dramatically decreases the rate of charge-transfer quenching; In addition to Type II fragmentation, the a-thio- alkoxyacetophenones undergo competitive B-cleavage ix: Michael Jeffrey Lindstrom varying degrees. The extent of B-cleavage is found to depend upon the nature of the leaving group. For instance, 2-methylsulfinylacetophenone reacts exclusively via B-cleavage, whereas for 2-thiomethylacetophenone such cleav- age is a minor pathway. The competition between y-hydrogen abstraction and B-cleavage was separated and its effect upon kct determined. The rate data obtained through the standard Stern- Volmer analysis were separated into inductive and resonance constituents. The following order, listed in decreasing ability to stabilize an adjacent radical center, was deter- mined: OMe > SPh ~ SBu ~ OPh > OH > 80311 > Ph > Me > SOZMe. PART II Phenacylsulfides A number of substituted phenacylsulfides were synthe- sized and studied to further investigate the nature of B-cleavage. All compounds studied underwent photoinduced B—cleavage in benzene with the concomitant production of the appropriate acetophenones and radical coupling products. The addition of 0.05M benzenethiol was observed to maximize the quantum yields and eliminate the formation of out-of- cage coupling products. The rate constants of B-cleavage were determined by standard Stern-Volmer analysis. Electron donating substitu- ents on the benzene ring decrease the rate of B-cleavage Michael Jeffrey Lindstrom significantly. The rate of B-cleavage is also dependent on the nature of the leaving group. Relative rates of B-cleavage for various groups were determined as follows: SPh > SOMe > StBu > SOzMe. 6-Substituted Phenyl Ketones A variety of 5-substituted phenyl ketones were syn- thesized and studied. All compounds formed varying amounts of 4-benzoyl-l-butene in addition to acetophenone upon uv irradiation. The effect of solvent upon the ratio of Type II products to 4-benzoyl-l-butene is negligible, indicating that elimination of HX was occurring by a free radical pathway rather than ionically. The mechanism was determined to proceed by initial 1,4 diradical formation followed by B-cleavage and rapid in-cage disproportionation to yield 4-benzoyl-l-butene and HX. Comparison of Type II products to 4-benzoyl-l-butene ratios allowed calculation of relative rates of B-cleavage for a variety of groups. A novel cyclic phenyl ketone system in which the 6-substituents and benzoyl groups are inaccessible to each other was synthesized. Smooth photolytic elimination of HX in these cases was observed. Afairly large rate enhancement for y-hydrogen abstraction was observed for trans-4-bromo- l,4-dimethyl-l-benzoylcyclohexane and is attributed to anchimeric assistance by bromine. Competitive a-cleavage in these compounds revealed information concerning ground state equilibria. To Joan p. 5.4. ACKNOWLEDGMENTS The author wishes to thank Professor Peter J. Wagner for his inspiring guidance throughout the course of this endeavor. His friendship, advice, and sense of humor have made the years of "courses and requirements" at M.S.U. very enjoyable. It has indeed been a privilege and a pleasure to work with him. The author would also like to thank the Chemistry Department at M.S.U. for financial support and the use of its excellent facilities. Thanks also to the NSF for research assistantships administered by Dr. Wagner. Very special thanks is extended to my parents, grandparents, relatives, and friends for their support and encouragement. My wife, Joan, deserves the author's deepest appre- ciation for her patience, companionship, and support. Her editorial comments, advice, and typing of the manuscript proved invaluable. iii TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . LIST OF FIGURES . . . . . . . . . . LIST OF SCHEMES . . . . . . . . . . INTRODUCTION . . . . . . . . . . . Objectives and Organization . . . The Norrish Type II Reaction . . Charge-Transfer . . . . . . . . . Energy Transfer . . . . . . . . . The Norrish Type I Reaction, a-Cleavage B-Cleavage . . . . - . . . . . . Stern-Volmer Kinetics . . . . . . PART I. INTRAMOLECULAR CHARGE-TRANSFER QUENCHING OF EXCITED STATES BY SULFUR . . . Results . . . . . . . . . . . . . Preparation and Product Identification . . . Quantum Yields . . . . . . . . Quenching Studies . . . . . . Intersystem Crossing Yields . Discussion . . . . . . . . . . . General Observations . . . . . Rate of Hydrogen Abstraction Versus Inductive and Resonance Effects . . . iv Page xi xv xvii 12 15 17 17 17 18 l9 19 30 3O 32 Page Competitive B-Cleavage and Y-Hydrogen Abstraction . . . . . . . . . . . . . . . . 38 Charge-Transfer Quenching . . . . . . . . . . . 4l Conformational Effects . . . . . . . . . . . 42 Effect of Donor on kct . . . . . . . . . . . 45 Bimolecular Quenching . . . . . . . . . . . 48 Indications for Further Research . . . . . . . 49 Synthetic Utility . . . . . . . . . . . . . 49 Products from C-T State . . . . . . . . . . 50 Effect of Ring Substituents . . . . . . . . 50 Nature of Quenching by the Sulfonyl Group . . 50 Energy Transfer Resulting in Homolytic Cleavage . . . . . . . . . . . . . . . . 51 PART II. RADICAL B-CLEAVAGE VIA EXCITED STATES AND PHOTOGENERATED DIRADICALS . . . . . . . . . . 52 Phenacylsulfides (Results) . . . . . . . . . . . . 52 Synthesis and Identification of Photoproducts . . 52 Quantum Yields . . . . . . . . . . . . . . . . 53 Quenching Studies . . . . . . . . . . . . . . . 53 Intersystem Crossing Yields . . . . . . . . . . 54 Spectroscopic Studies . . . . . . . . . . . . . 54 Phenacylsulfides (Discussion) . . . . . . . . . . 68 Mechanism . . . . . . . . . . . . . . . . . . . 68 Quantum Yields . . . . . . . . . . . . . . . . 72 Spectroscopy . . . . . . . . . . . . . . . . . 74 Rates of B-Cleavage and Charge-Transfer . . . . 79 Indications for Further Research . . . Estimation of In-Cage Coupling . . Generation of Free Radicals . . . . 6-Substituted Phenyl Ketones (Results) . Synthesis . . . . . . . . . . . . . . Identification of Diastereomers . . . Identification of Photoproducts . . . Quantum Yields . . . . . . . . . . . . Quenching Studies . . . . . . . . . . 6-Substituted Phenyl Ketones (Discussion) Mechanism of Radical B-Cleavage via Photogenerated Diradicals . . . . . Quantum Yields and Other Observations Rates of B-Elimination . . . . . . . . y-Hydrogen Abstraction and Anchimeric Assistance . . . . . . . . . . . . Conformational Effects . . . . . . . . Indications for Further Research . . . Synthetic Utility . . . . . . . . . Relative Rates of 8-Cleavage . . . B-Substituted Butyrophenones . . . Inductive Effects . . . . . . . . . EXPERIMENTAL . . . . . . . . . . . . . . . . Preparation and Purification of Materials Solvents and Additives . . . . . . . . Benzene . . . . . . . . . . . . . . Methanol . . . . . . . . . . . . . vi Page 84 84 85 85 85 87 89 91 91 101 101 106 107 112 114 119 119 119 120 120 121 121 121 121 121 Dioxane . . . . Pyridine . . . . Acetonitrile . Ethanol . . . . Hexane . . . . . Pentane . . . . Benzenethiol . . Internal Standards Dodecane . . . Tetradecane . Hexadecane . . Heptadecane . Octadecane . . . Nonadecane . . . Quenchers . . . . . Napthalene . . . 1-Methylnapthalene . cis- and trans-1,3-Pentadiene cis-1,3-Pentadiene n-Butylsulfide . .n-Butylsulfoxide n-Butylsulfone . Ketones . . . . . . Acetophenone . . Valerophenone Butyrophenone . vii Page 121 121 121 122 122- 122 122 122 122 122 122 122 122 122 122 122 122 123 123 123 123 123 123 123 123 123 Phenacylchloride . . . . . . . . . . Phenacylbromide . . . . . . . . . . . B-Bromopropiophenone . . . . . . . . y-Chlorobutyrophenone . . . . . . . . 6—Chlorovalerophenone . . . . . . . s-Chlorohexanophenone . . . . . . . . 2-Thiomethylacetophenone . . . . . . 2-Thiobutylacetophenone . . . . . . . 2-Thio-t-butylacetophenone . . . . . 2-Thiophenylpropiophenone . . . . . 2-Thiophenylisobutyrophenone . . . . Phenyldesylsulfide . . . . . . . . . 2-Thiophenylacetophenone . . . . . . 2-Thiophenyl-4'-f1uoroacetophenone . 2-Thiophenyl-4'-chloroacetophenone . 2-Thiophenyl-4'-bromoacetophenone . . 2-Thiophenyl-4'-methylacetophenone . 2-Thiophenyl-4'-methoxyacetophenone . 2-Thiopheny1-4'-thiomethylacetophenone 2-Thiophenyl-4'-cyanoacetophenone . 2-Thiophenyl-4'-pheny1acetophenone . 2-Thiophenyl—4'-dimethylaminoacetophenone l-Benzoyl-l-thiophenylcyclopropane 4'-Thiophenylmethylacetophenone . . . 3-Thiobuty1propiophenone . . . . . 4-Thiobutylbutyrophenone . . . . . . viii Page 124 124 124 124 124 124 125 125 125 125 125 125 125 125 125 126 126 126 126 126 126 126 127 127 128 128 4-Thio-t-butylbutyrophenone . 4-Thiophenylbutyrophenone . 5-Thiobutylvalerophenone . . 5-Thio-t-butylvalerophenone . 5-Thiophenylvalerophenone . 6-Thiobutylhexanophenone . . 3-Butylsulfinylpropiophenone 4-Butylsulfinylbutyrophenone 5-Butylsulfinylvalerophenone 5-Phenylsulfinylvalerophenone 3—Butylsulfonylpropiophenone 4—Butylsulfonylbutyrophenone 5-Buty1sulfonylvalerophenone 5-Phenylsulfonylvalerophenone 2-Methylsulfinylacetophenone 2-Methylsulfonylacetophenone 5-Thioacetoxyvalerophenone . 5-Thiocyanatova1erophenone . trans-4-Chloro-l,4-dimethyl-1-benzoyl- cyclohexane . . . . . . . cis- and trans-4-Bromo-1,4-dimethyl-l- benzoylcyclohexane . . . . 4'-Methoxy-5-chlorovalerophenone 4'-Trifluoromethyl-5-chlorovalerophenone 5-Iodovalerophenone . . . . . 5—(4'-Methoxy)-phenoxyvalerophenone . 4-Benzhydryloxybutyrophenone ix Page 128 129 129 129 129 129 129 130 130 130 130 131 131 131 131 131 132 132 132 133 134 134 134 134 135 Page Techniques . . . . . . . . . . . . . . . . . . . 135 Preparation of Samples . . . . . . . . . . . . 135 Photochemical Glassware . . . . . . . . . . 135 Stock Solutions and Photolysis Solutions . . 136 Quantum Yields, Quenching, and Sensiti- zation Studies . . . . . . . . . . . . . 136 Degassing . . . . . . . . . . . . . . . . . 136 Irradiation Procedures . . . . . . . . . . . . 137 Kinetic Runs . . . . . . . . . . . . .°. . 137 Preparative Runs . . . . . . . . . . . . . 137 Analysis of Samples . . . . . . . . . . . . . 137 Identification of Photoproducts . . . . . . 137 Gas Chromatography Procedures . . . . L . . 140 Actinometry and Quantum Yields . . . . . . 141 Spectra . . . . . . . . . . . . . . . . . . . 143 LIST OF REFERENCES . . . . . . . . . . . . . . . . . 147 APPENDIX 0 O O O O O O O O O O O O O O O O O O O O O 158 LI ST OF TABLES Table Page 1. Photokinetic Data for Various Ketosulfides . . . 21 2 . Photokinetic Data for Various Ketosulfoxides . . . 24 3. Photokinetic Data for Various Ketosulfones . . . 25 4. Bimolecular Rate Constants for Quenching by Sulfur . . . . . . . . . . . . . . . . . . 26 5. Calculated Rate Data for Sulfur Containing Ketones O O O O O O O O O O O O O O O O O O 0 3 3 6. Relative Resonance Stabilization Factors for Various Groups . . . . . . . . . . . . . 34 -7. Approximate Amounts of B—Cleavage Versus y-Hydrogen Abstraction in a-Thioalkoxy- acetophenones . . . . . . . . . . . . . . . . 40 8. Calculated Rate Data for a-Thioalkoxyaceto- phenones . . . . . . . . . . . . . . . . . . 40 9. Quantum Yields and Rate Data for p-Substituted Phenacylsulfides . . . . . . . 55 10. Quantum Yields and Rate Data for Structurally Variant B-Benzoylsulfides . . . . . . . . . . 56 'k * ll. Substituent Effects on n,n and n,w Transi- tions in Phenacylsulfides . . . . . . . . . . 58 12. Triplet Energies of Ring-Substituted Phenacylsulfides . . . . . . . . . . . . . . 59 13. Approximate Percentage of IBP and MAP Found in the Photolysis of t-ZSPh . . . . . . . . . 74 14. Calculated Rate Data for Various Phenacyl- sulfides . . . . . . . . . . . . . . . . . . 80 15. Relative Rates of B-Cleavage in Phenacyl- sulfides . . . . . . . . . . . . . . . . . . 83 xi Table 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Quantum Yields of Acetophenone and 4-Benzoy1- 1-Butene Formation from Various G-Substituted Valerophenones . . . Solvent Effects on B-Cleavage for Various S-Chlorovalerophenones . . . . . . Quantum Yields and Rate Data for 4-Halo-l,4- Dimethyl-l-Benzoylcyclohexanes . .' Rate Constants of B-Elimination of Various Substituents in G-Substituted Valero- phenones O O O O O I O C I O O I 0 Rate Constants of y-Hydrogen Abstraction and a-Cleavage for the 4-Halo-l,4-Dimethyl-l- Benzoylcyclohexanes . . . . . . . Ground State Equilibria of 4-Halo-1,4- Dimethyl-l-Benzoylcyclohexanes . . Data for 2-Thiomethylacetophenone Data Data Data Data Data Data Data Data Data Data Data Data Data Data for for for for for for for for for for for for for for Thio—t-butylacetophenone 2-Thiobutylacetophenone . 4-Thiobuty1butyrophenone . . 4-Thio-t-butylbutyrophenone 4-Thiophenylbutyrophenone . 5-Thiobutylva1erophenone . . 5-Thiophenylvalerophenone . 6-Thiobuty1hexanophenone . 5-Thiocyanatova1erophenone 5-Thioacetoxyvalerophenone . 2-Methylsulfinylacetophenone 4-Butylsulfinylbutyrophenone 5-Buty1sulfinylvalerophenone 2-Butylsulfonylacetophenone xii Page 93 94 95 108 115 118 160 160 161 161 162 162 163 163 164 164 165 165 166 166 167 Table 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. Data Data Data Data Data Data Data Data Data for for for for for for for for for 5-Butylsulfonylbutyrophenone . . 5-Butylsulfonylvalerophenone . . . . 2-Thiophenylacetophenone . . . . . . 2-Thiophenyl-4'-F1uoroacetophenone . Thiophenyl-4'-chloroacetophenone . . 2-Thiopheny1-4'-bromoacetophenone . 2-Thiophenyl-4'-methy1acetophenone . 2-Thiophenyl-4'-methoxyacetophenone 2-Thiophenyl-4'-thiomethoxyaceto- phenone . . . . . . . . . . . . . . . . . Data for 2-Thiophenyl-4'-dimethy1aminoaceto- phenone . . . . . . . . . . . . . . . . . Data Data Data Data Data Data Data Data Data Data Solvent Effects on Quantum Yield for 5-Chloro- for for for for for for for for for for 2-Thiophenyl-4'-phenylacetophenone . 2-Thi0phenyl-4'-cyanoacetophenone . 2-Thiophenylpropriophenone . . . . . 2-Thiopheny1isobutyrophenone . . . l-Thiophenyl-l-benzoylcyclopropane . 4'-Thiophenylmethylacetophenone . . 4'-Thiophenylacetophenone . . . . . Phenyldesylsulfide . . . . . . . . . 5-Phenylsulfinylvalerophenone . . . 5-Phenylsulfonylvalerophenone . . . valerophenone . . . . . . . . . . . . . . Solvent Effects on Quantum Yield for 5-Chloro- 4'-methoxyvaler0phenone . . . . . . . . . Solvent Effects on Quantum Yield for 5-Chloro- 4'-trifluoromethylvalerophenone . . . . xiii Page 167 168 168 169 169 170 170 171 171 172 172 173 173 174 174 175 175 176 176 177 177 178 178 Table Page 60. Data for trans-4-Chloro-l,4-dimethyl-l- benzoylcyclohexane . . . . . . . . . . . . . 179 61. Data for trans-4-Bromo-l,4-dimethyl-l- benzoylcyclohexane . . . . . . . . . . . . . 180 62. Data for cis-4-Bromo-l,4-dimethyl-l- benzoylcyclohexane . . . . . . . . . . . . . 181 63. The 3-Thiobutylpropiophenone Sensitized Isomerization of cis-l,3-pentadiene . . . . 181 64. The 4-Thiobutylbutyrophenone Sensitized Isomerization of cis-l,3-pentadiene . . . . 182 65. The 3-Butylsulfinylpropiophenone Sensitized Isomerization of cis-1,3-pentadiene . . . . 182 66. The 3-Butylsulfonylpropiophenone Sensitized Isomerization of cis-l,3-pentadiene . . . . 182 xiv LIST OF FIGURES Figure l. Stern-Volmer Quenching Plot for Acetophenone Formation from 4-SBu (O) , 4-StBu (.) , and 4-SPh ([5) in Benzene with 1,3- pentadiene . . . . . . . . . . . . . . . . 2. Quenching of Acetophenone Formation from Butyro- phenone with BuSBu (A) , BuSOBu (.) , and BuSOzBu ([3); Quenching of p-MeO Valero- phenone with BuSOBu (A) and BuSOzBu (Q) . . 3. Dependence of uantum Yield for 3-SBu (C)) and 4-SBu ( ) Sensitized Isomerization of cis-1,3-pentadiene on Diene Concen- tration in Benzene . . . . . . . . . . . . 4. Hammet Plot of Relative Rates of Triplet State Y-Hydrogen Abstraction for G-Substituted ValerOphenones. (Points Cl and CN are included for comparison and can be found in reference 76.) . . . . . . . . . . . . . 5. Quantum Yield of Acetophenone Formation in 2-SPh Versus [¢SH] . . . . . . . . . . . . 6. Stern-Volmer Quenching Plot for the Quenching of the p-Substituted Acetophenone Forma- tion from Cl-ZSPh (O), Br-ZSPh (Q), Me-ZSPh (A), and CN-ZSPh (A) with Napthalene in Benzene . . . . . . . . . . . 7. Ultraviolet Spectra of 2-SBu. [A = 0 . 00183M(EtOH) , B = 0. 0014M(heptane) , C = 0. 000183M(EtOH) , D==0.00014M (heptane)] . . . . . . . . . . 8 . Phosphorescence Spectrum of 2-SBu in MTHF at 77K . 9 . Phosphorescence Spectrum of 2-SPh in MTHF at 77K . lO. Phosphorescence Spectrum of OMe-ZSPh in MTHF at 77K 0 O O O O O O O O O O O O O O O O O XV Page 27 28 29 37 60 61 62 63 64 65 7 Figure Page 11. Phosphorescence Spectrum of NMez-ZSPh in MTHF at 77K . . . . . . . . . . . . . . 66 12. Phosphorescence Spectrum of CN-ZSPh in MTHF at 77K 0 I O O O O O O O l O O O O O O 67 13. X-Ray Crystal Structure of trans-CDMBC . . . . 96 14. NMR Spectra of trans—CDMBC in CDCl (Bottom); in cnc1 Containing 0.0259g/o.5%1 Eu(fod)§(Top) . . . . . . . . . . . . . . 97 15. NMR Spectra of trans-BDMBC in CDCl (Bottom); in cuc1 Containing 0.0259g/o.531 Eu(fod)g(Top) . . . . . . . . . . . . . . 98 16. NMR Spectra of cis-BDMBC in CDCl3 (Bottom); in CDCl Containing 0.0259g/0.5ml Eu(fod)§(Top) . . . . . . . . . . . . . . . 99 17. Temperature Dependence of qu for Valero- phenone (‘) and 6-Iodovalerophenone (A) in Benzene . . . . . . . . . . . . . . . . 100 18. 13C Spectrum of trans-CDMBC in CHCl3 . . . . . 144 19. 13C Spectrum of trans-BDMBC in CHCl3 . . . . . 145 20. 13C Spectrum of cis-BDMBC in CHCl 146 3 . . . . . . xvi Scheme 1. LIST OF SCHEMES The Mechanism of the Norrish Type II Reaction 0 O O O O O I O O O O O O O O 0 Available Pathways for the Biradical Formed from a-Cleavage . . . . . . . . . . . . Photolytic Cleavage of Pivalophenone . . . Photochemical Fate of l-Methyl-l-Benzoyl- cyclohexane . . . . . . . . . . . . . Mechanism of B-Cleavage of B-Ketosulfides . Addition of Thiols to Vinylcyclopropane . . Synthesis of 4-Halo-l,4-Dimethyl-l- Benzoylcyclohexanes . . . . . . . . . . Photolysis of 5-Thiophenoxyvalerophenone in Benzene . . . . . . . . . . . . . . . . Conformational Effects in the Photochemistry of trans-CDMBC . . . . . . . . . . . . . xvii Page 10 10 12 68 73 86 105 116 INTRODUCTION Objectives and Organization The original objective of this thesis was, in a broad sense, to investigate structure-reactivity relationships in organic photochemistry. In particular, the effect of chain length upon the rate of intramolecular charge-transfer quenching of the excited benzoyl group by various sulfur moieties was to be investigated. However, as work progressed, unexpected results provided the stimuli to stray away from the original objective and pursue other paths. Thus, this thesis is composed of two separate but related parts. In an attempt to avoid duplication, the introduction is composedcnf one part in which topics relevant to the problems investi- gated will be discussed. The results for each part are treated separately and the corresponding discussion follows each section. The experimental section presents the proce- dures and materials for both parts. Part I deals solely with the effect of chain length upon the rate of quenching of the benzoyl group by various sulfur moieties. In particular, ketosulfides, sulfoxides, and sulfones were investigated. Part II deals with radical B-cleavage of halo and thiyl radicals both from excited states and photogenerated biradicals. This includes the photochemistry of phenacyl sulfides, 5-halo enui 5-thiyl substituted valerophenones, and a novel cyclic system. The Norrish Type II Reaction Ketones possessing y-hydrogens undergo a charac- teristic 1,5-hydrogen transfer, followed by cleavage or cyclization, upon uv irradiation. Norrish and Appleyard1 first observed this reaction in 1934 while investigating the photodecomposition of methyl butyl ketone in the gas phase. Yang,2 who first noted the formationxxfcyclobutanols as products, hypothesized the intermediacycxfa 1,4 biradical. Compelling evidence for the intermediacy of a 1,4 biradical 4 In fact, the bi- has been presented by Wagner3 and others. radical from y-methoxybutyrophenone has actually been trapped by thiols.5 The formation of cleavage products and of cyclobutanols is collectively called the Norrish Type II Reaction. Scheme 1 presents the accepted mechanism. For phenyl alkyl ketones the rate constant of inter- 11 -1 system crossing, k isr~10 sec ;so fluorescence, kf, isc’ which occurs on the order of 106sec-l, and radiationless decay, kd, are negligible. As a result, the intersystem crossing yield is equal to unity in most cases.6 Thus,1flua triplet manifold is responsible for the majority of the photoreactivity observed in phenyl ketones. Population of R J Scheme 1. The Mechanism of the Norrish Type II Reaction. the triplet manifold results in y-hydrogen abstraction followed by either reverse hydrogen transfer, k_Y,cn:product formation, kcyc and k3. Reverse hydrogen transfer was first postulated by Hammond and Wagner7 to account for low quantum yields i3: systems where the excited state reactivity was thought to be fairly high. Wagner8 later showed that reverse hydrogen transfer could be eliminated by the use of polar solvents or various additives that are good Lewis bases. Phosphorescence, k from the triplet is slow relative to p’ y-hydrogen abstraction and is usually insignificant. The reactivity of the triplet benzoyl group has been likened by Walling and Gibian9 to that of the t-butoxy radical. Numerous studies investigating structure-reactivity relationships have been presented by Wagnerlo 11 and others, but in general the variation in reactivity in phenyl ketones can be ascribed to either inductive or resonance effects on the y-radical center or the nature of the excited triplet. Electron releasing substituents on the benzene ring generally lead in) decreased or negligible reactivity of the triplet benzoyl due to the interposition of a 3n,n* state, which is unreactive toward y-hydrogen abstraction.12 Interestingly, Wagnerl3 has shown that thermal equilibration of the 3n,n* state with the higher 3n,n* state in p-methoxy- valerophenone is taking place. In this case, the low equi- librium concentration of the upper 3n,n* state produces a low observed rate constant for y-hydrogen abstraction. Little work concerning the effect of structure on cleavage:cyclobutanol ratios has been done, but Lewis14 has shown that a-substitution on phenyl alkyl ketones markedly 15 increases the amount of cyclization. Wagner has noted an analogous effect for a-fluoroketones. Charge-Transfer The phenomenon of charge-transfer was first invoked by Mulliken16 to explain why iodine is violet in CCl but 4 brown in benzene. Formally, one can envision charge-transfer as simply partial electron transfer from a donor to an acceptor. D°°+ A' -—+ D'+A‘(C-T complex) Photochemical precedent was established by Leonhardt 17 and Weller who demonstrated by flash spectroscopy that solutions of perylene and an amine undergo electron transfer, 18 later a process not possible in the ground state. Cohen observed that benzophenone isphotoreduced by Et3N 1000 times faster than by isopropanol. This led to the suggestion that rapid electron transfer from the nitrogen to the triplet carbonyl followed by proton transfer was responsible for the 19 later suggested the observation. Davidson and Lambert following mechanism for the photoreduction of benzophenone by amines. + - - 2NCHZR) -—> R COH + RCHNR * u o _ R c=o 3 + RCH NR ——a (ch-o R 2 2 2 2 2 Later experiments by Cohen and Chao20 definitively ruled out direct hydrogen abstraction from the amine. The ionization potentials of amines have been correlated with their ability to interact with triplet ketones. Cohen and 21 Guttenplan noticed an inverse relationship between the ionization potential of the amines and their abilitytxnphoto— reduce a ketone. However, they noticed the results were inconsistent with full electron transfer and could best be described as partial electron transfer. The fact that small 6 solvent effects were noted, relative to those for ground state electron transfer reactions, further supported their contention. Wagner and Kemppainen22 also noted no increase in the rate of quenching of valophenone by tri- ethylamine or t-butylamine in CHBCN relative to that in benzene. Other types of heteroatoms have demonstrated the ability to form C-T complexes with excited carbonyls. Thio- ethers have been shown to quench the phosphorescence of benzophenone and also its photoreduction by isoborneol.23 Interestingly, little photoreduction by sulfides possessing a-hydrogens was found to occur, indicating that proton trans- fer analogous to that observed for amines24 is not an important pathway for sulfides. The 4-carboxybenzophenone sensitized photooxidation of methionine was found to occur by initial charge-transfer complexation between triplet benzophenone and sulfur followed by internal electron transfer to nitrogen.25 Phosphorous, antimony, and arsenic have also been postulated to quench the Type II Reaction of butyrophenone via a charge-transfer interaction.26 Intramolecular charge-transfer quenching has not been extensively investigated. However, Wagner27 has found that intersystem crossing yields decrease and rates of quenching increase as an amino group is moved closer to the carbonyl. These facts indicate a competition between intersystem crossing, charge-transfer quenching, and y-hydrogen abstrac- tion. Type II products have also been shown to arise from the C-T state of y-dimethylaminobutyro-2-napthone by ‘protonation of the ketyl radical anion.28 Strict conformational requirements for C-T quenching were demonstrated by Wagner and Scheve.29 The lack of C-T quenching in 1:3 demonstrates that such quenching requires through-space orbital overlap. Ph [.1 I (D Intramolecular charge-transfer quenching of excited ketones with sulfur has not been previously measured. However, a "C-T type" intermediate has been proposed for the photorearrangement of 2 to 3.30 <35$®~©é~<§ A variety of optically active sulfoxides have been shown to undergo photochemically induced stereomutation, 31 resulting in racemization. The charge-transfer nature of these transformations is indicated by the fact that energy transfer would be at least 10-15 kcal endothermic.32 Sulfones have not been previously shown to interact with excited carbonyls, but the conversion of 4 to 5 hints at some sort of direct interaction.33 63;“ V I v o 802 r1-—L—_> Pry/H h i .5. Energy Transfer Other mechanisms besides C-T interactions are possible for the quenching of triplets. When an acceptor has a lower triplet energy than the donor, triplet-triplet energy transfer occurs. Other types of energy transfer are possible, but in the case of phenyl ketones, intersystem crossing yields are usually high so triplet-triplet energy transfer predominates. Usually, the effectiveness of a quencher is determined by the position of its lowest triplet level and not by its molec- ular structure.34 On the other hand, the rate of exothermic triplet energy transfer in solution is dependent on the viscosity of the solvent, which for moderately viscous solvents is described by the modified Debye Equation.35 k = k et diff = 8RT/2000n However, in less viscous solvents the quenching rate constant is still lower than the calculated diffusion rate. The implication of this is that there is inefficiency in the energy transfer which may be indicative of a preferred configuration of the donor and acceptor molecules. Interestingly, not much is known about the steric requirement to promote the exchange interaction needed for transfer; however, Wagner and McGrath36 have shown that transfer is rapid at van der Waals separation of donor and 37 acceptor. Cowan and Baum have measured ke in styryl- t ketones where the separation between donor and acceptor varies from 2 to 4methylene groups. Predictably, when 11 M-1 -1 n = 2, ke =~lO sec . As n is increased to 4, t 1 -1 k = ~109 M- sec . Thus, the number of conformations in et which the two ends of the molecule interact decreases rapidly as the number of methylene groups is increased. It has been estimated that for every 1.23 increase in the distance between the donor and acceptor, an order 38 of magnitude decrease in ket results. The Norrish Type I Reaction, a-Cleavage The Norrish Type I Reaction involves homolytic cleavage of the 1,2 bond in cyclic and acyclic ketones, resulting in the formation of an acyl and alkyl radical pair. Many other classes of compounds, such as the carboxylic acid derivatives,39 which possess a heteroatom as the a-substituent, also undergo this reaction. Thus, the reaction will be referred to, in a broader sense, as a-cleavage. The fate of the radical pair varies depending largely on structural 10 40 but the three possible routes as pictured in features, Scheme 2 are recoupling, two modes of disproportionation, and decarbonylation. R H/CM . hv C ROH /\/\/fi\0 -————9 ‘__9W/”\v/A\C:CJO'_—€> R §_____ \ Scheme 2. Available Pathways for the Biradical Formed from a-Cleavage CC) + ,/”\»/”\3 In phenyl ketones, a-cleavage leads to a slightly different picture than that presented in Scheme 2. As shown below in Scheme 3, the photolysis of pivalophenone41 results in the formation of a benzoyl radical and a t-butyl radical. In-cage radical-radical recombination competes effectively with \\:SH Ph/ji\H + ,/'_' coupling Ph/I products + H_<_ Scheme 3. Photolytic Cleavage of Pivalophenone. ll diffusion out of the cage. Disproportionation to form benzaldehyde and isobutylene can occur in-cage or out-of—cage. However, in the presence of thiols, which are efficient 42 radical scavengers, the intermediate radicals are trapped, thus minimizing or eliminating out-of—cage disproportionation and coupling. While a-cleavage is the major mode of decomposition 43 for certain aliphatic ketones, it is usually not observed for normal phenyl alkyl ketones. Even in a,a-dimethylvalero- phenone, Y-hydrogen abstraction is sufficiently fast that a—cleavage comprises only about 5% of the reactivity.44 In fact, triplet aliphatic t-butyl ketones a-cleave about 4000 45 times faster than triplet pivalophenone. The nature of the excited state has also been found to greatly influence the 46 rate of a-cleavage. Lewis has shown that pivalophenone, * which has a lowest 3n,n state, cleaves with a rate constant of around 107'sec-l; but p-phenylpivalophenone, which has * a 3n,w lowest triplet, is essentially stable to photolysis. Lewis47 has noted a competition between a-cleavage and y-hydrogen abstraction in some very elegant work with several cycloalkyl phenyl ketones. Scheme 4 presents the conforma- tional possibilities for 1-methyl-l-benzoylcyclohexane (6) -Interestingly, 6:3 only undergoes y-hydrogen abstraction, whereas 6-e exhibits only a-cleavage. In this case, kII and kI were faster than interconversion of the excited conformers, so product ratios reflected the ground state population of the respective conformers. Thus, Lewis determined that the ground state equilibrium was about 65:35 in favor of 6-e. 12 Type II products Type I products A A) kII kI 3 o 3 o Kax Keq At It ‘c ‘,R ___9 P 4___ Ph 6-a 6-e Scheme 4. Photochemical Fate of l—Methyl-l-Benzoylcyclo- hexane. B-Cleavage There is a wide variety of photochemical reactions of carbonyl moieties in which electronic excitation results in initial cleavage of one of the bonds 8 to the carbonyl group.48 Also typical are B-eliminations in free-radical 49 reactions. In fact, a-haloacetophenones and, to a lesser extent, phenacylsulfides have been used to photoinitiate 50 polymerization reactions. To date no kinetic studies on B-cleavage reactions of ketones have been done. 13 There are, however, several interesting photoreactions involving B-elimination. The photoisomerization of . . . . 5 isothiochroman-4-one IS a c1a531c example. 1 We... 0‘, W O . l o // .<———— co;— Surprisingly, when the benzene ring is replaced with a napthalene, the reaction does not go. The photolysis of 1 yields 8 by B-cleavage of oSPh followed by disproportionation and aromatization.52 0 ¢ hv -S¢ O |\l (D \l + OH 00 H89?) 8 [CD 14 53,54 In other systems, where disproportionation is not possible, coupling products are observed as shown below. 0| R L '0' H + sulfur containing Ph/b\/s\ PM?) prOdUCtS o v o /°\/S‘>2’r “ > AVA + i—som” The formation of the vinylsulfonate in the second example arises from radical recombination on oxygen rathertflunicarbon as shown below. [Asa/R] + - ---e + 15 Stern-Volmer Kinetics The quantum yield for any photochemical process can be expressed as a product of probabilities. Thus, for the Type II Reaction: in = c) (1) 1. ?-k+kd (2) where ¢isc equals the intersystem crossing yield, kY equals the rate constant of hydrogen abstraction, kd equals the rate constant of triplet decay, T equals the triplet lifetime, and P equals the probability that the diradical will go on to product. In the presence of an external quencher, GD becomes: = kY + kd + kq[Q] (3) all-4 where kq equals the bimolecular rate constant. Utilizing these facts, the Stern-Volmer Equation (4)55 can be derived. 0 = 1 + qu [Q] (4) 9+9 In the presence of polar solvents or Lewis bases, which minimize reverse hydrogen transfer, the maximum quantum yield can be expressed: (5) Thus, k = (6) 16 The sensitization equation (7)56 allows measurement Of ¢isc° -1 -1 -1 1 sens. - isc ktr[Q]) (7) Thus, in the case of 1,3-pentadiene, a equals 0.55 and a plot of 0.55/q)c+t versus [cis-l,3-pentadiene]-l yields a straight line in which the reciprocal of the intercept equals ¢isc and intercept divided by the slope equals ktTr° PART I INTRAMOLECULAR CHARGE-TRANSFER QUENCHING OF EXCITED STATES BY SULFUR ‘ Results Preparation and Product Identification A variety of ketosulfides of varying chain length were prepared by Sn2 displacement on an appropriate haloketone by the corresponding sodium thiolate. The ketosulfoxides and sulfones were prepared by hydrogen peroxide oxidation of the corresponding ketosulfides. The corresponding compounds and their abbreviations are listed in Tables 1, 2, and 3. For example, a-thiomethylacetephenone is denoted as 2-SMe and a-methylsulfinylacetophenone as 2-SOMe. All compounds except 3-SBu, 3-SOBu, and 3-SOzBu pro- duced acetophenone upon uv irradiation. These compounds were stable to irradiation at 31303 and underwent no disappearance of ketone. Acetophenone was actually isolated by preparative vpc for 4-SBu but was subsequently identified either by the appearance of a singlet at 2.36 in the nmr or by comparison of its vPc retention time with that of an authentic sample. The olefinic moiety which results from cleavage was not observed since it probably came out under the solvent peak on the vpc. Small peaks which were assumed to be the 17 18 cyclobutanols were sometimes observed and usually corres- ponded to about 10-15% of the total product. No evidence of the corresponding thietanols was observed for 2-SMe and 2-SBu. The 6-substituted valerophenones were found to produce varying amounts of 4-benzoy1-l-butene (4-BB) in addition to acetophenone. The mechanism for the formation of 4-BB will be discussed in Part II. Quantum Yields Quantum yields for acetophenone formation and other products were determined by irradiation at 3130A in a merry- go-round apparatus at 25°C. Solutions containing 0.05M ketone in benzene, which sometimes contained various addi- tives, were irradiated in parallel with degassed benzene solutions of 0.1M valerophenone, which served as an acti- nometer.57 All samples were degassed by two or three freeze- thaw cycles prior to irradiation. Percent conversion of the ketone and/or actinometer was kept below 10% whenever possible. Disappearance quantum yields were measured at typically 20-30% conversion. Product to standard ratios were measured by vpc. The associated error was estimatedtnrdupli- cate runs and was generally found to be about i5% for aceto- phenone formation and about :10-15% for disappearance quantum yields. Polar solvents or Lewis base additives had little or no effect upon the quantum yields. The maximum quantum 19 yields, ¢ I were usually obtained in benzene containing max 1.0M dioxane. The quantum yields are listed in Tables 1, 2, and 3. Quenching Studies ‘Stern-Volmer quenching slopes were performed at 3130A or 36603 by photolysis of 0.05M ketone solutions containing varying amounts of either 1,3-pentadienecnzl-methylnapthalene. Conversions were usually kept below 10% for the tube with no quencher, and the slopes were linear outtx>a ¢°/¢ value of about 7 or 8. The associated error was estimated by dupli- cate runs to be about i10%. Values obtained from these studies are listed on Tables 1, 2, and 3. A representative Stern-Volmer plot is presented in Figure 1. The bimolecular rate constants for quenching by n-butylsulfide, n-butylsulfoxide, and n-butylsulfone were determined by an analogous procedure through the utilization of butyrophenone and p-methoxybutyrophenone as substrates. No quenching of acetophenone formation was detected when n-butylsulfone was used as a quencher. The bimolecular rate constants are listed in Table 4, and the quenching plots are presented in Figure 2. Intersystem Crossing Yields Intersystem crossing yields were determined by parallel irradiation at 31303 of 0.05M ketone solutions containing varying amounts of cis-1,3-pentadiene and a 0.10M 20 acetophenone solution containing 1.0M cis-l,3-pentadiene, 58 Plots of reciprocal the latter serving as an actinometer. quencher concentrations were linear. In all cases where measurement was possible, ¢. was equa11x31. Irreproducible isc results were obtained for the a-thioalkylacetophenone and the G-substituted valerophenones due probably to the production of low concentrations of radicals. The intersystem crossing yields are presented in Tables 1, 2, and 3, and a representa- tive sensitization plot is presented in Figure 3. 21 Anmumuec %m>\/U\nm m~.m me.H o.H s~.o uuu nu- a Asmmuec mo.m m.eoa.e amm\\/(\)/AV\:E o~.o n oe.a o.H He.o n mH.e mH.o H~.o a Asmmumc smm//\\/4v\nn mom.o n OH.H o.H In- In- nu- m Asmmnmc umo.o n mm.o emm\//o\HE H.o n e.H nu- mo.o n me.o me.o so.o n me.o a Asmumumc Jx/m\\//o\xrw v.0 n m.m nu: umoo.o n eo.o Hoo.ov nu- : o Amzmnmv omo.o n mm.o mzmx\/xw\en ~.o n eo.m In- mm.o In- He.o a cox one 539 e VT e 9509.50 3 m .mmowwdumoumx m50aum> How sumo owumcflxouonm .H manna. 22 .ommumv Oflm//\\/(\// m Redo.o _ + as us. mas o mm o nu- o Azomsmc m N . 0 20m <m .po>uomno mcoumx mcfluumum mo mocmumommomflp ozm .ocmucon :H ocwpwuhm 2o.Hm .ucm>aom Hocmnumzo .ocmncon ca mme 2mo.oo .mcoucon cw ocmxoflp So.an .MOMHM Um QCGNGQQ C.“ GCOCGSQOUQUM HO COHHMEHOH 05v... HON mmVHmflxn SUCMSOM A.e.ueooc .H manna 24 .mcmusnlanahoNcmnlvo .mcaeensn zom.oe .mme ch.oo .mcmxoflp Zo.Hn .momam um ocmncon cw poumwpmuuHm AsmOmumc mmm.o 9m _ m o.nmo.o H mm.o owl/\)/(\)/J\LL ~.o n N.H~ in- nme.c mo.o mm.o r Asmomuev emmo.o smom\/\\/o\HE om.o n e.mH o.H nmo.o mo.o us: a Asmomnmv o~.o n oe.a o.H nu- In: In- a Asmuomumc . - . . prom\\//J\irm nu- nu- oeo o + mm o ma 0 u)- w Amzomumc mzom>o\;m no.0 n ma.o u)- ome.o n ee.o HH.o u)- a v one xms ml p x e e Q.Me e panomfioo .mmpflxOMHSmoumm msoflum> MOM sumo owumcflxouonm .m manna 25 .ocmusnlalahoNconlv @ .mcmeMp 2o.Ho .mcmucmn :H mme Smo.o Q .Momam um mcmucmnfijncwumflomuuHm Annmom-mc N am On e.o~oo.o n mo.e /\/<\/o\§H me n mom I.) one.o n oe.o nu: me.o : o Asmmomnec - sm~0m\\/(\)/AW\:E ems + omsm o H oom o uuu nu: a Aammomumv . In- In- nu- , _ was o H o omumv nu- nu- has e me o u-) r AsmNOmnmv I DMN0m>U\Sm as + saw I-) nem.o om.o nu- : 0 sex ones xmee we sue easoneoo .mmc0mazmoumx msowum> MOM mama owuocwxouonm .m manna 26 Table 4. Bimolecular Rate Constants for Quenching by Sulfur. Quencher kq’lOSMIlsec-l n-butylsulfide 3.25a n-butylsulfoxide 0.014a 0.012 c n-butylsulfone --- aMeasured by quenching the formation of acetophenone from butyrophenone. bMeasured by quenching the formation of acetophenone from p-methoxyvalerophenone. cNo quenching of acetophenone formation from butyrophenone or p-methoxyvalerophenone observed. 27 6‘ 1k 5.. 4‘ A. ¢O I 3- A. 2 ‘ (5 6 0 1. 0.5 1.0 1.5 2.0 [l,3-pentadiene],M Figure 1. Stern-Volmer Quenching Plot for Acetophenone Formation from 4-SBu (O), 4-StBu (O), and 4-SPh(A) in Benzene with 1,3-pentadiene. e|e Figure 2. 28 )4 A» .A ea A9 2%— 0.5 1.0 1.5 2.0 [Q] .M Quenching of Acetophenone Formation from Butyro- phenone with BuSBu (A), BuSOBu (7.) , and BuSOZBu(Z§);and Quenching of p-MeO Valerophenone with BuSOBu (‘) and BuSOZBu(@) . 29 Figure 3. *O *0 O '3 —l I r I I I 0.5 1.0 1.5 2.0 2.5 3.0 [cis-pentadiene]"':",M"l Dependence of Quantum Yield for 3-SBu (D) and 4-SBu (.) Sensitized Isomerization of cis-l,3- pentadiene on Diene Concentration in Benzene. 30 Discussion General Observations As indicated in Tables 1, 2, and 3, the quantum effi- ciencies for acetophenone formation are fairly low, typically 5 around 20-30%. Normally, in the absence of alternative decay routes the inefficiency in the Type II Reaction is due to 9 In the revertibility (reverse 1,5-hydrogen transfer).5 present system, however, only a modest increase in quantum yield is observed in the presence of additives (e.gu,jpyridine or dioxane) which normally maximize the Type II yield. For example, 4-SBu goes from a quantum yield of 0.13 in benzene to only 0.18 in benzene containing 1.0M dioxane. Generally, enhancements by a factor of 2-3 are observed. Thus, intra- molecular hydrogen bonding between the ketyl and sulfur moieties must be faster than reverse hydrogen transfer H-bonding >> k-x) ' H Ph’g;v;T/SR k . PW’Cf . H-bonduEL \ o *7 ' \ hv Type II E:\\N #:2é: products 0 R M P Such intramolecular hydrogen bonding has been postulated previ- 60 (i.e., k ously to explain the elimination of ROH in B-alkoxyketones . 31 Difficulty in maximizing the quantum yield has been noted previously for 6-methoxyvalerophenone and was ascribed in part to 6-hydrogen abstraction.61 The propensitycflfsulfides, sulfoxides, and sulfones toward hydrogen bonding is well documented;62 and as such, intramolecular hydrogen bonding must be minimizing revertibility in the same manner as external Lewis bases.63 Polar solvents seem to depress cyclization slightly since solvation of the hydroxyl moiety presents some steric inhibition to coupling.64 In contrast, analogous steric problems are absent for intramolecular solvation; and, in fact, an increase in cyclization might be expected since the conformation necessary for intramolecular hydrogen bonding favors that necessary for coupling. However, material balances indicate that the normal amount of cyclization is taking place, i.e., lO-20%. The anomalous lack of any thietanol formation for 2-SMe and 2-SBu has been noted 65 . . . and seems surprising Since the analogous com- previously pounds containing oxygen or nitrogen cyclize with ease.66 Ground state reversion of the initially formed thietanol to starting material can be accommodated by two pathways; and as such, isomerization of the starting materials would be expected as shown below. H % '9 2 88 Ph 8 ——)ph/C\/S\/RP S__9PP/C\‘/\ 32 Since isomerization of starting material was not observed,67 this process can be eliminated. Therefore, it seems likely the differing behavior of this system with respect to the analogous oxygen and nitrogen systems may be related to the relative weakness of the C-S bond (65 kcal) as compared to the C-0 (85 kcal) and C-N (73 kcal) bonds. Also, the interposition of the fairly large sulfur atom may introduce subtle conformational effects, which inhibit cyclization. Rate of Hydrogen Abstraction Versus Inductive and Resonance Effects Calculation of the various rate constants deserves comment. The value of ky, the rate constant for y-hydrogen abstraction, can be calculated from the measured lifetime and the maximum quantum yield of Type II products according to equation 6. k = (6) The calculated rate data is listed in Table 5. Now, in the case of the a-thioalkoxyacetophenones, B-cleavage is found to compete with y-hydrogen abstraction, so in these cases the observed kY actually equals kY + k The extenttx>which B' B-cleavage competes in the thioalkoxyketones and the corres- ponding corrections will be discussed in the following section. The rate constant of C-T quenching will be discussed in the appropriate section. 33 Table 5. Calculated Rate Data for Sulfur Containing Ketones. Compound l,1085ec-l k lO-lsec-l k ,107sec-l T Y’ ct 2-SMe 24.3 ---a 149 2-SBu 31.3 —--a 135 Z-StBu 12.8 ---a 118 2-SOMe 51.0 ---a 61.0 2-SOZMe 0.23 —--a 1.4 3-SBu 45.4 --— 455 3-SOBu 29.4 --- 294 3-SOZBu 0.28 --- 2.84 4-SBu- 29.4 52.9 241 4-SOBu 3.21 0.96 31.0 4-SOZBu 0.013 0.03 0.10 4-StBu 28.9 78.0 211 4-SPh 10.5 37.8 67.2 S-SBu 1.80 3.96 14.0 S-SOBu 2.36 9.91 13.7 5-sozau 0.24 1.01 1.42 S-SPh 1.30 3.89 9.09 6-SBu 1.37 3.42 10.2 S-SCN 0.47 1.2 3.5 S-SAc 1.14 9.15 2.28 aSee Table 7 for the separation of kY + k B. 34 All substituents stabilize a radical center relative to the unsubstituted case. The fact that the various kY values vary greatly with substitution indicates that inductive deac- tivation of the y-hydrogen center must be strong. In the case of y-substitution, separationxmfthe competing inductive 68 has and resonance effects is at best tricky; but Wagner separated the two by assuming a p value of -4.3 for y-substituents. Table 6 lists various relative stabilization factors which were calculated using the data in Table 5. Table 6. Relative Resonance Stabilization Factors for Various Groups. Substituent OI 10-4'3OI k/ko Stagiiiggtion SPh 0.30 0.051 3.0 58.8 SBu 0.25 0.084 4.5 53.6 SOBu 0.52 0.0058 0.077 13.3 SOzBu 0.60 0.0026 0.022 0.78 OMe 0.30 0.051 5.0 98.0 Me --- --- 1.0 1.0 OH 0.25 0.085 3.1 37.0 0 0.10 0.37 3.1 8.4 O¢ 0.39 0.021 1.15 54.7 Table 6 reveals the following sequence listed in decreasing ability to stablize an adjacent radical center: OMe > SPh -SBu ~'O¢ > OH > SOBu > 0 > Me > 502Me 35 As with alkoxy radicals,‘59 the stabilizing effect of y-substitution reflects the availability of the lone pair of electrons. Ethers are about 2-5 times better than the corresponding alcohols at stabilizing an adjacent radical center. Qualitatively, it is known that sulfur can stabilize a free radical since the a-hydrogens of sulfides are suscep- tible to hydrogen abstraction.70 The present study indicates a factor of about 2 less than the corresponding ether. Interestingly, SBu and SPh stabilize an adjacent radical to about the same extent, indicating a lack of participation by the benzene ring. Although phenyl systems are known to participate in free radical stabilization,71 ESR studies have indicated no participation by the phenyl ring when attached to a heteroatom.72 The stabilizing effect of a methyl group is fairly small relative to heteroatoms as one might expect. The difference in kY for a primary carbon versus a secondary carbon, as in butyrophenone versus valerophenone, is about 16. However, a difference of about 2-0 between primary and secondary is apparent for 2-SMe and 2-SBu. In the case of y-methoxybutyrophenone versus y-methoxyvalerophenone, the 73 former is slightly more reactive. Here the greater reac- tivity of a 3° relative to 2° carbon is not great enough to offset the larger number of hydrogens in the latter. The same effect has been noted for the a-alkoxyacetophenones,74 where the ky's for a-methoxy-auxin-ethoxyacetophones are 9 l 9 3.2- 10 sec- and 8.4 -10 sec-l, respectively--a factor of 2.6. 36 Introduction of an insulating methyl group as in the d-substituted valerophenones corresponds to a decrease in the inductive effect by a factor of 0.4375 and at the same time eliminates any resonance stabilization of the y-radical center. Figure 4 presents a Hammettgflrnzfor 6-substituents in which a plot of log(k/k°), where ko equals the rate constant of y-hydrogen abstraction for valerophenone, versus 76 0 yields a p-value of -l.85. Wagner has calculated an I identical p—value for various 6-substituents, and agreement here indicates the correctness of the present values and further illustrates the predictive value of such a plot. However, an anomalous result arises in that 5-SOBu falls way off the line and appears that y-hydrogen abstractionifisabout ten times faster than that expected on purely inductive grounds. An analogous activation has been noted for 6-Br and 6-I valerophenone and was attributed to anchimeric assistance in the hydrogen abstraction step.77 Comparable rate enhancements have been observed previously for sul- I fides,78 and Shevlin79 has recently observed a weak anchi- meric effect by a sulfinyl group. Whereas assistance by bromine or iodine and formation of a bridge species can be pictured below, 37 0.0‘ A (8080) -0.5' k log-é; k r -1.0- 0.2 0.4 0.6 0.8 “’1 Figure 4. Hammet Plot of Relative Rates of Triplet State y-Hydrogen Abstraction for 6-Substituted Valero- phenones. (Points C1 and CN are included for comparison and can be found in reference 76 . ) 38 assistance by a sulfinyl group can be represented in two ways: 0 R R ‘\S/’ . .41 °r 1.... Unfortunately, only speculation concerning the nature of anchimeric assistance by sulfinyl groups is possible with the available data. The topic of anchimeric assistance will be discussed in depth in Part II. Competitive B-Cleavage and y-Hydrogen Abstraction For the B-ketosulfides, sulfoxides, and sulfones, B-cleavage is found to compete with y-hydrogen abstraction. This reaction is well documented80 but has escaped a thorough kinetic examination. Thus, 2-SBu has two available routes for acetophenone formation as shown below. diffusion 1) out of cage 9 R [ p \. 'SBU ] 2)RSH 'Ph/o\+ BUSH fl kBZ/‘SB h Ph/CVS\/\/ ‘b k H \Y 6 3 Type II P h/éVSW products 81 82 Lewis and Wagner have separately estimated that about:50% of the initially formed radicals in Type [cleavage recombine in-cage to form starting material as indicated. Thus, 39 compounds which form acetophenone primarily via B-cleavage experience at least a 50% inefficiency. The addition of thiols has been shown to increase the quantum yield of product formation by trapping the radicals that diffuse out of the cage.83 Enhancements on the order of two to five are typical in the present system as well as previous systems.84 Examination of Tables 1, 2, and 3 allows estimation of the amount of B-cleavage relative to y-hydrogen abstrac- tion in the thioalkoxyacetophenones. For instance, 2-StBu, which possesses no y-hydrogens, forms acetophenone in the presence of 0.05M HSPh with a quantum yield of 0.04. Thus, at least 4% of acetophenone formation arises from B-cleavage in 2-SMe and 2-SBu. However, if one assumes about 50% in- cage recombination, 8% of the light is attributable to B-cleavage. If the total quantum yield of acetophenone is divided into that found via B-cleavage, the competition between B-cleavage and Y—hydrogen abstraction can then be separated. The results are presented in Tables 7 and 8. Interestingly, 2-SOMe and 2—SO Me appear to react mainly via 2 B-cleavage, which is in agreement with that observed by 85 86 Majeti and deMayo. A more detailed discussion of B-cleavage appears in Part II. 40 Table 7. Approximate Amounts of B-Cleavage Versus Y-Hydrogen Abstraction in a-Thioalkoxyacetophenones. Com ound % B-Cleavage % y-Hydrogen p Abstraction 2-SMe 21 79 2-SBu 14 86 2-SOMe 100 0 2-SOzMe 83 17 Table 8. Calculated Rate Data for a-Thioalkoxyacetophenones. a 8 -1 8 -l 8 -1 Compound ¢corr kY,10 s k8,10 s kct'lo s 2-SMe 0.39 7.5 1.9 14.9 2-SBu 0.57 15.3 2.49 13.5 2-StBu 0.08 --- 1.02 11.8 2-SOMe 0.88 --- 44.9 6.1 2-502Me 0.41 0.016 0.078 0.14 aQuantum yield of acetophenone corrected flar50% recombination of the initially formed radicals. 41 Charge-Transfer Quenching Previous work on aminoketones has indicated a competi- 'thx1betweenintersystem crossing, charge-transfer quenching, and y-hydrogen abstraction. Since ki *- 10 sec for so 87 phenyl ketones, interaction of the amine with the excited carbonyl must be comparably rapid to be able to quench the singlet. However, intersystem crossing yields for the ketosulfides are found to be one, indicating a much slower rate of interaction. Despite this, the general trends should be the same--that is, as the sulfur moiety is moved closer to the carbonyl, the amount of charge-transfer quenching should increase. The amount of charge-transfer quenchingiJ1 the present case is assumed to be the only significant decay process of the triplet, as expressed by the following equation: 1. -‘F_kY+kCt (8) Thus, knowledge of 'l' and the quantum yield allows calculation of kct’ The effect of competing B-cleavage on kCt has already been discussed in the previous section. Examination of Table 5 reveals fairly high kct values for 5-SBu and 6-SBu at 7 1 7 14.2-10 sec" and 10.2-10 sec'l, respectively. As previously found in azidoketones,88 kct is found to decrease by a factor «ii/'3 Ph’Ph/gvsvph —’ PWWSVPh . Ph’P ' W l_0 / TYPE 11 PRODUCTS (1) = 0.04 45 inert to hydrogen abstraction. Perhaps coincidentally, the quantum yield Padwa observed was identical to the value measured for 2rStBu in the presence of 0.05M thiol.T%erefore, it is possible that products arise in Padwa's system via B-cleavage followed by hydrogen abstraction from starting material which possesses highly abstractable hydrogen atoms as shown below. 9 10 h" > h/Ph/C\. + 'Svph —‘ 7 P /19 O B-cleavage n P. PH ph’ - Unfortunately, the available data ch) not allow further speculation. Product formation from the charge-transfer state in the present system has not been rigorously investigated. Effect of Donor on kct- In the case of phenylamines, the nitrogen leads to increased electron densitijlthe ring which, in turn, leads to a more favorable interaction between excited species and the conjugated system, resulting in more efficient quenching relative to aliphatic amines.96 However, phenylsulfides are less efficient quenchers than aliphatic 97 sulfides, which indicates that quenching arises primarily 46 by interaction of the non-bonding electrons of sulfur. Electron withdrawing groups on the sulfur, like phenyl, decrease its effectiveness. Thus, 3-SPh quenches about 98 noted a difference of four times slower than 3-SBu. Cohen about nine between n-butylsulfide and phenylsulfide in the bimolecular quenching of the photoreduction of benzophenone by isoborneol with sulfides. That S-SAc and S-SCN quench about seven times slower than 5-SBu is also indicative of the effect electron withdrawing groups have on the avail- ability of the non-bonding pairs on sulfur. Expectedly, increasing the oxidation state of the sulfur moiety leads to a decrease in the rate of charge- transfer quenching. For instance, 4-SOBu quenches about eight times more slowly than 4-SBu, and 4-SOZBu about 2500 times more slowly than 4-SBu. The reduced rates of quenching for sulfoxides relative to sulfides is attributable to the decreased availablity of the electrons on sulfoxides. This, in turn, is manifested in a higher ionization potential than sulfides.99 The nature of the C-T complex in sulfides and sulfoxides is pictured below and can be represented as interaction of the non-bonding electrons on sulfur with the triplet. , O 6 18R 5 C?- JLSR (DJ/(R Ufa l Pn/4%\c_/J Ph/€K\__/J f%y/¢ p , 47 Sulfones, on the other hand, are completely oxidized and possess no non-bonding electrons except for thosecn1oxygen; and, as such, the previous representation seems inappropriate. Sulfones can best be represented by a resonating decet 100 structure. R O- R O_' R O \ x“ ES<1 é—-————9 /}£;f e———___9 >34' R \o R 0‘ R/\o‘ Formally, C-T quenching can arise by interaction with either the lone pair of electrons on oxygen (lla) or with a n-bond between the sulfur and oxygen (11b). . \ IO ' O - (I)- + S haamusuoauum How sumo mumm can mvamww Esusmso .OH OHQMB 57 .Uonsmmme DOZM .ea magma :e 0000efl 000 00H0es 2500000 0000000060 .mu 60 .manmzocmsqcso mms Zmo on .Momam um mcmucmn ch 5mm I vao.o woo.o Amomv sm\L// __ 0 cm - See -..- 30003 Jaw me e n 9 _Me vasomfioo 1.0.ucoo. .oa 0Hnme 58 'k * Table 11. Substituent Effects on n,n and 0,0 Transitions in Phenacylsulfides.a * * Compound xmax(n’" ), nm Amax("," ),nm 3 340 (372) 240 (8,157)}3 say/C‘\/’SB“ 336 (634)b 242 (10,382) 8 /K::ZTA3\V/Sph X e H 341 (448) 246 (19,462) CN 355 (809) 248 (31,181) OMe 336 (649) 257 (19,603) SMe 340C 307 (20,183) NMe2 330c 325 (27,067) aIn heptane; molar extinction coefficients in parentheses. * bIn ethanol; fine structure on n,n band has disappeared. * * C:10,“ buried under «,1: ; estimated by comparison of fine struc- ture with other compounds. 59 Table 12. Triplet Energies of Ring-Substituted Phenacyl- sulfides.a Compound Et(kcal) H 70.9 (74) C Bu SPh X x H 73.5 (74) CN 68.4 (69.2)b OMe 70.4 (71) NMe2 62.6 aNumbers in parentheses refer to the corresponding ring- substituted valerophenone in 2-methylpentane. (Seeref.112) bE. J. Siebert, unpublished. 60 0.3 ~ 0 j 4 0.2 0 0.1 - S i i I fij r I 0.1 0.2 0.3 0.8 [48H] Figure 5. Quantum Yield of Acetophenone Formation in 2-SPh Versus [05H]. 61 6.4 5‘ 4-1 2° . ¢ A 3'1 0.5 1.0 1.5 2.0 [napthalene],M Figure 6. Stern-Volmer Quenching Plot for the Quenching of the p-Substituted Acetophenone Formation from Cl-ZSPh (O), Br-ZSPh (Q), Me-ZSPh (A), and CN-28Ph.(‘k) with Napthalene in Benzene. 62 .10:0060:120Hooo.o u a .Amoumvzmmaoco.o n o .Amcmummnvzeaoo.o n m .Amoumvzmmaoo.o n ma .smmlm mo mnuommm umaoa>muuaa .5 musmflm Ezomv cow omm mNm oom mhm omm com 63 350 400 450 500 550nm Figure 8. Phosphorescence Spectrum of 2-SBu in MTHF at 77K. 64 350 400 450 500 550nm ‘ Figure 9. Phosphorescence Spectrum of 2-SPh in MTHF at 77K. 65 350 400 450 500 550nm Figure 10. Phosphorescence Spectrum of OMe-ZSPh in MTHF at 77K. 66 350 400 450 500 550nm Figure 11. Phosphorescence Spectrum of NMe -28Ph in MTHF at 77K. 2 67 370 420 470 520 570nm Figure 12. Phosphorescence Spectrum of CN-ZSPh in MTHF at 77K. 68 Discussion Mechanism The free radical nature of photoinduced B-cleavage of B-ketosulfides has been firmly established by careful examination of product distributions.113 On the other hand, certain substituents display a propensity toward heterolytic cleavage as in the case of a-tosylketones, which result in rearranged products typical of carbonium ion chemistry.114 In the present system, the detection of coupling products-- namely, diphenyldisulfide--and the fact that thiols, which are known radical scavengers, greatly enhanced the quantum yields, eliminated heterolytic cleavage as a mechanistic possibility. The mechanism of photoinduced B-cleavage of B-ketosulfides is presented in Scheme 5. diffusion out ofcmge RSH O coupfing (sh/[L~ products Scheme 5. Mechanism of B-Cleavage of B-Ketosulfides. 69 As in free radical systems,115 fairly strict stereo- chemical alignments between the radical center and the bond that is breaking would be expected. Thus, two important conformations need to be considered: conformation 13! in which the bond that is breaking is parallel to the benzoyl fl-system, and conformation 13, in which the breaking bond is parallel to the C-0 bond and the sulfur and oxygen are eclipsed. Previous studies on carbonyl compounds116 containing a-heteroatoms have indicated that conformations 12 and 13 are present in about equal amounts in solution. In fact, d—haloaldehydes and ketones exist primarily in conformation _]._3_ where the carbonyl and halogen atom are nearly eclipsed. However, in the case of a-thioalkoxyaldehydes117 conformation 12 is clearly dominant. The analogous thio- alkoxyketones have not been studied. On the other hand, IR J: studies have shown a-bromocyclohexanone exists primarily in the conformation where the bromine is axial, while the a-chlorocyclohexanone exists in the conformation where it is equatorial. The situation is obviously fairly complicated; however, other factors aside, conformation 12 clearly represents the sterically favored orientation. In fact, such a conformation seems rigidly required by analogy with 70 119 120 the a-tosyloxy- and a-aminocyclohexanones which cleave only when axial. Further support for conformational control of B-cleavage can be gleaned from the fact that phenyldesyl- sulfide (PDS) reacts with a quantum efficiency far below the other phenacylsulfides. Examination of the three impor- tant conformational possibilities reveals a possible reason for this. SPh 16 I“ .b In this case, alignment of the proper orbitals as in 14 and 15 causes serious eclipsing interaction between the two phenyl rings or the carbonyl and phenyl ring. Therefore, the amount of conformation 16 present at any time will be higher than normal; and as such, the amount of C—T quenching will be increased. Consideration of conformation E does not suggest straightforward type of elimination. However,the geometry enables maximum overlap of the n-orbital on oxygen and one of the lone pairs on sulfur. This conformation allows C-T complexa- tion as discussed in Part I and product formation from such a complex must be considered. One possible mode of cleavage can be envisioned as follows: 71 However, this seems unlikely in light of the fact that an analogous mechanism is also possible for equatorial a-tosyloxy- and a-aminocyclohexanones, which do not react. Also, 2-SOtBu cleaves smoothly in spite of the fact that SOtBu would be a poor anionic leaving group.121 O n/\- + -SO+ Pmiy; + ’80—}— Another mechanistic possibility exists. Direct energy transfer from the carbonyl to the sulfur moiety could result in homolytic cleavage as shown below. b I '0' ‘ ? Ph/k/sph ——+ Ph/VSPh ——> PRODUCTS The preferred orientation necessary for energy transfer in this case is unclear. Charge-transfer complexation, on the other hand, between the n-orbital and the lone electron pair 72 on sulfur can occur only from conformation 13. Triplet- triplet energy transfer for 2-SPh would be slightly endo- thermic since the triplet energy of thioanisole is ~76 1:001.122 As substitution on the benzoyl moiety lowers the triplet energy, the exchange would become even more endothermic. In any event, products from homolytic cleavage in 3-SPh and 4-SPh were not observed; so it seems that product formation via energy transfer is unlikely. Therefore, conformation 13 results mainly in C-T quenching of the excited carbonyl. The amount of this quenching will be discussed later. Quantum Yields Tables 9 and 10 present the quantum yields for a variety of structurally variant phenacylsulfides. The addi- tion of benzenethiol leads to fairly large quantum yield enhancements. Out-of—cage coupling is a fairly efficient process, usually leading to respectable yields of dibenzoyl- 123 ethane. Near quantitative material balances in the presence of thiols indicate that all the radicals are scavenged before they get a chance to couple. The rate constant for trapping by alkylthiols (kt) is<31the order of 107M-lsec-l.124 The rate constant for trapping by PhSH.can be estimated from the addition of thiols to Vinylcyclopro— pane125 as shown in Scheme 6. When benzenethiol is used, 100% of product E is formed, which indicates that kt > kc. If kcnv107sec-l,126 kt for PhSH must bel~108M-lsec-l. RSH %_1_7 %_8_ Rs?- CH SH 12 88 \A 08H 100 0 lg Scheme 6. Addition of Thiols to Vinylcyclopropane. Examination of Table 10 reveals some interesting information regarding s-ZSPh and t-ZSPh. Although the dis- proportionation products--i.e., vinylacetophenone and a-methylacrylophenone--could not be separated from their saturated analogues under the analytical conditions, nmr experiments on t-ZSPh allowed approximate measurement of the amount of a—methylacrylophenone (MAP). Curiously, no enhancement of quantum yield was observed for either compound upon the addition of PhSH. Table 13 lists the approximate amounts of reduction and disproportionation products from t-SPh as estimated by nmr experiments. In the presence of thiol, it is assumed that most of the radicals that diffuse out of the cage are trapped. Thus, most of the dispropor- tionation product (MAP) in the presence of thiol is probably coming from in-cage disproportionation. The slight 74 Table 13. Approximate Percentage of IBP and MAP Found:h1the Photolysis of t-ZSPh. fl Solvent Phj\< PM ¢IPB + MAP (IBP) (MAP) Benzene 45 55 0.33 Benzene 83 17 0.34 0.05M 05H predominance of MAP in the absence of thiol is also indica- tive of in-cage disproportionation. The amountscflfIBP and MAP are comparable in the absence of benzenethiol,indicating that out-of—cage disproportionation greatly predominates over coupling as is the case for t—butyl radicals.127 Data arexufi:availab1e for s-ZSPh, but one would expect a lesser predominance of disproportionation over recombination due to the diminished amount of B-hydrogens.128 Spectroscopy Any discussion of substituent effects upon the rate]'LA transition of acylbenzenes,they tend to destabilize n,0* transitions, eventually resulting in inversion of the triplet levels for strong donors.132 The effect of substituents upon the * * ordering of n,0 and 0,0 states is presented below. 0,0 * n,0 \ * 0,0 * nzn CH3 CF3 OCH3 H SCH3 Table 12 presents the triplet energies of some ring- substituted phenacylsulfides. The phosphorescence spectrum of 2-SBu is presented in Figure 8 and strongly resembles valero- phenone in which the vibrational structure corresponds to the ground state carbonyl stretch.133 The triplet energy is about 3 kcal lower than valerophenone due to the inductive effect of the a-thioalkoxy group, which stablizes the n,0* state slightly. The spectra of the a-thiophenyl- acetophenones (Figures 9 through 12), on the other hand, * show a marked lack of structure, resembling more 0,0 states. 76 Disturbingly, the E for the a—thiophenylacetophenones are t almost identical to the corresponding ring-substituted valerophenones. The discrepancy here is why does SBu lower the Et energy relative to the corresponding valero- phenone but SPh seems to have no effect. Several possibili- ties seem to exist. One possibility is that emission is being observed from two different n,0* triplets, analogous to that observed in the dual phosphorescence of phenylalkyl- ketones.134 Thus, two excited state conformations, the higher in energy of which resembles the ground state con- formation, would yield different Et's. Extrapolation to the phenacylsulfide case presents the possibility that 70.9 kcal represents the conformationally relaxed emission of 2-SBu, whereas the values for the a—thiophenylacetophenones repre- sent emission from a higher, unrelaxed state. Therefore, the increase in energy would offset the decrease expected on inductive grounds. The difference between SBu and SPh apart from extended conjugation in SPh are primarily steric in nature. Since dual phosphorescence is highly dependent upon the nature of the solvent,135 which in turn determines the orientation of the carbonyl and sulfur groups in the glassy matrix, subtle changes in conformational orientation could result. A more obvious possibility is that 2-SPh is actually a lowest 0,0* state (p-OMe and p-NMe2 are most certainly so), * whereas 2-SBu is n,0 in nature. The inductive effect of * the a-substituent would not be expected to affect a 0,0 at state as much as an n,0 state. 77 Lastly, a disturbing possibility is that B—cleavage to yield a radical pair is competitive with emission in the glassy matrix. If that is the case, the emission would actually be from the acetophenone moiety, and the lack of structure in the phosphorescence spectra could be explained by its free radical nature. The effect of a-substituents on the uv spectra is also interesting. The intensification of the n,0* transi- tion and its subsequent red shift relative to the unsubsti- tuted ketone has been noted previously for a-heteroketone spectra.136 In fact, the original observation by Jones137 has since been extensively used in conformational analysis. Again, two conformations need to be considered: one, where the C-X bond is parallel to the 0-system which corresponds to an axial substituent for a cyclohexanone as in conforma- tion 12, and two, where the C-X bond is parallel to the C-0 bond corresponding roughly to the analogous equatorial cyclohexanone substitutent as in conformation 13. The intensification of the n,0* absorption and its shift to longer wavelength is observed only for axial sub- stituents. Equatorial substituents on a-substituted cyclo- hexanones produce either no effect or cause a slight blue shift. Allinger138 has ascribed this effect to hyperconju— gation between the C-X bond and the carbonyl 0-system. Interaction between the carbonyl 0* level and a low lying C-Xo* orbital when X is axial gives rise to a red shift. Hoffmann139 has recently treated the a-aminoketones 78 theoretically and found Allinger's arguments essentially correct in that the hyperconjugation effect is extremely dependent upon C-N/0-system alignment. Little beyond thisis known about the (as-heteroatom effect, but the nature of the 140 effect can perhaps be likened to the Murrel or Labhart- 141 models for B-Y unsaturated carbonyls. In any Wagnier event, it seems that enhanced absorptions are observed when- ever a carbonyl and group of low ionization potential are present with proper mutual geometric disposition. Table 11 presents the uv data for selected substi- tuted phenacylsulfides, and Figure 7 shows the actual spectra of 2-SBu in heptane and in ethanol. As mentioned previously, fine structure is observed in the nnr* shoulder for all compounds in heptane but disappear when ethanol is used as solvent. This observation is a rare occurrencelnn:has been observed previously in cyclobutanone142 and exo-dicyclopenta- 143 dienone. Interestingly, the spacing of the structure corresponds to the stretching frequency of the excited carbonyl singlet (about 1200 cm-l).144 Differing behavior between a-SBu and a-SPh substi- tuents are also noted in the uv. SBu stabilizes the n,v* transition relative to acetophenone (Amaxn,fl* = 316nm) but does nothing to the 0,0* transition. However, SPh stabilizes both n,0* and 0,0* transitions. There is obviously some interaction between the phenyl ring and the benzoyl fl-system but the exact nature of this observation is unknown. As 145 noted previously, electron donating groups tend to 'k * stabilize the 0," transition while raising the n,0 79 * transition energy. Conversely, CN-ZSPh stabilizes the n,0 * transition but also lowers the 0,0 band because of its ability to extend the conjugation. Rates of B-Cleavage and Charge-Transfer The rate constant of B-cleavage, k is calculated 8! from equation 9. _ corr (9) where ¢co equals the observed quantum yield multiplied by rr two to correct for about 50% in-cage recombination. The remaining inefficiency is assumed to be due to C-T quenching and is calculated as mentioned previously from equation 8. The rate data are listed in Table 14. The rate of B-cleavage can be influenced by three factors: the nature of the excited state, the stability of the incipient radicals, and the overall thermodynamics of the system. A consideration of the effect of excited state upon reactivity centers upon the dichotomy between the n,0* and 0,0* states. As can be seen from the phosphorescence studies, all substituents tend to lower the triplet energy. However, the lowering of the triplet energy is accompanied by an opposing effect. That is, electron donors are indeed lowering the Et' but they are also inverting the * * ordering of the n,0 and 0,0 states. Incomplete knowledge 80 Table 14. Calculated Rate Data for Various Phenacylsulfides. Compound 0;:ira 31131085.l k8,1083-l ct.1083 2-SPh 0.48 --- >100 --- F-ZSPh 0.48 --- >100 --- Cl-ZSPh 0.48 62.4 30.0 32.5 Br-ZSPh 0.40 25.0 10.0 15.0 Me-ZSPh 0.60 --- >100 --- OMe-ZSPh 0.82 22.7 18.4 4.30 SMe-ZSPh 0.38 1.67 0.63 1.04 NMez-ZSPh 0.38 3.62 1.38 2.24 0-23Ph 0.54 1.25 0.68 0.58 CN-ZSPh 0.32 46.3 14.8 31.5 s-ZSPh 0.64 -—- >100 --- t-ZSPh 0.68 --- >100 --- cy-ZSPh 0.84 36.2 30.4 5.80 4'-MSPh 0.80 --- >100 --- 2-StBu 0.08 12.8 1.02 11.8 4'-SPh <0.002 --- --- -—- PDS 0.02 --- --- --- aCalculated by doubling the observed quantum yield. 81 of the multiplicity of the reaction prevents a good corre- 1ation between the phosphorescence and uv data and the observed reactivity. However, all reactivity comes from the triplet state in the compounds which are quenchable--it is the efficienty of triplet formation which is unknown. One general observation is applicable--i.e., electron donors tend to slow down the reaction. However, this immedicately implies that 0,0* states are less reactive than n,0* states. The reduced hydrogen abstraction ability of the carbonyl in the 30,0* state is understandable in terms of its greatly reduced electrophilicity,146 but the mechanism of B-cleavage requires only that there be free spin on the carbonyl carbon. The valence bond representation of the respective states are shown below. 57’ .11. n,0 We * The main difference between the two is that in the n,0 state, an electron from the n-orbital on oxygen is promoted to a 0*-orbital, whereas the 0,0* state results from promo- tion of an electron from the 0-system to a 0*-orbital. Both representations, however, indicate significant amounts of spin on the carbonyl carbon, but in actuality it is known * that most of the excitation in 0,0 states is localized in 82 the benzene ring.147 Thus, it would appear that the observed decrease in rate with electron donating substituents is due to 21 decreased amount of spin density at the carbonyl carbon. In simpler terms, the n,0* state resembles a 1,2- biradical, whereas a 0,0* does not. Table 14 reveals a difference in the observed rate of C-T quenching, k between the electron donating and ct’ releasing substituents. Charge-transfer occurs primarily between the lone electron in the n-orbital and the lone pair on sulfur in the n,fl* state. Not much C-T formation between the 0-system in the 0,0* state and sulfur is expected since the oxygen is already "electron rich." Thus, as the lowest triplet becomes 0,0*, a decrease in kct is observed. Any<>4P involving the 0-system would in any event be decreased by electron donating groups. The effect of a p-CN-substituent is curious. It has the effect of decreasing the rate of cleavage much like electron donors; and, in fact, its phos- phorescence spectrum resembles a 0,0* state. However, its strong electron withdrawing effect increases the amount of kct by lowering the reduction potential of the carbonyl. As mentioned in the Introduction, the interposition of a 0,0* state for a n,0* state has been shown to eliminate (or greatly slow down) B-cleavage.148 As is the case in free radical chemistry, the sta- bility of the incipient radicals in part determines their rates 149 of formation. The kB for cy-ZSPh is less than t-ZSPh. The difference here presumably lies in the difference in 83 a tertiary radical and a 3° cyclopropyl radical. The instability of the cyclopropyl radical probably reflects its inability to attain a planar sp2 configuration.150 The cyclopropyl radical is surpassed only by the phenyl radical in difficulty of formation.151 The nature of the leaving group--i.e., the sulfur moiety--probably has the greatest effect on the rate of B-cleavage. Table 15 lists some relative rates100 >196 aData from Table 8. That -SPh is more stable than SR has been demonstrated 152 repeatedly. The order of stabilities--i.e. , SPh > SOR > SR > SOzR--confirms Rice‘s153 original suggestion that SOR has the greatest kinetic stability compared to SR or SO R. 2 A more detailed discussion of leaving groups will be forth- coming in the section on G-substituted valerophenones. 84 The strength of the C-S bond in the phenacylsulfides is intrinsically related to the stability of the incipient radicals. The bond dissociation energy for CH3SPh+ ~CH3 + 154 and would probably be lower ~SPh equals about 60 kcal/mole for 2-SPh since an d-keto radical is being formed rather than a methyl radical. Thus, even in the case of NMez-ZSPh, where the triplet energy is about 62 kcal/mole, the energy requirements are sufficiently met. However, as the triplet energies are lowered, the gap between the bond dissociation energy and the triplet energy becomes smaller, or less exothermic, and may explain the observed decrease in rate of B-cleavage. The bond dissociation energy for CH3SCH3 + ~CH3 + ~SCH3, which would correspond roughly to B-cleavage in 2-SMe, is about 73 kcal/mole. Indications for Further Research Estimation of In-Cage Coupling. A more accurate esti- mates for the amount of in-cage radical recombination could be gained by studying the following compounds: l+ Generation of the two radicals might be possiblekurextrusion of SO2 from the thiosulfonate, thus yielding the desired 85 radical pair. There is, however, the question astx>whether this would be an in-cage process. A measurement of the extent of racemization of an optically active phenacylsulfide might give some indication of the amount of in-cage recombination. Generation of Free Radicals. This system offers an excellent way to generate free radicals. Processes such as intramolecular or intermolecular additions could be studied. 0 II o O I SR H . 'SR hv etc. —> O? -—> CO —» ¢ // This system also offers an excellent opportunity to study solvent effects upon radical reactions since little is known in this respect. 6-Substituted Phenyl Ketones Results Synthesis The 6-substituted valerophenones listed in Table 16 were prepared by Sn2 displacement of d-chlorovalerophenone with the sodium salt of the appropriate nucleophile. To eliminate direct interaction between the benzoyl group and G-substituent as the cause of elimination of HX, compoundstI which the benzoyl group and X were inaccessible to each other ‘were synthesized as outlined in Scheme 7. 86 0 Nb \L__ Benzene 2 1 \/co Me > // § // 2 AICI3 Q 1)KOH 1 2)PhLi O O X =CL. «flNWBC) . he ”X 0 h X :8 BDMBC ‘ r ( ) Acetic acid Scheme 7. Synthesis of 4-Halo-1,4-Dimethyl-l-Benzoylcyclo- hexanes. The Diels-Alder Reaction between isoprene and a-methyl- methacrylate proceeded smoothly in benzene in the presence of a catalytic amount of aluminum chloride. Interestingly, a high degree of regioselectivity is observed in that only the 4-methyl isomer is isolated. The directing effect of AlCl3 and other Lewis acids is fortunate because the thermal reaction yields significant amounts of the 3-methyl isomer. Unfortunately, the analogous cycloaddition utilizing oi-methylacrylophenone and isoprene did not go. It was feared that thermolysis would lead to a inseparable mixture, so this path was not pursued further. Subsequent hydrolysis of the ester to yield the acid and reaction with two equiva- lents of PhLi to form the phenyl ketone proceeded in 95% and 60% yields, respectively. Addition of HX was found to be 87 exceedingly sluggish in ether or hydrocarbon solvent but was found to proceed smoothly in glacial acetic acid. In the case of the bromo-derivative, the two diastereomers were separable by fractional crystallization, but such was not the case for the Cl isomer for which only one diastereomer was obtained. Identification of Diastereomers Spectral comparison of the two bromide diastereomers revealed significant differencesiJithe chemical shift in the 1H and 13 C for the l-methyl group. Based on experimental free energy differences between axial and equatorial substi- 5 . . tuents,ls the folloWing assumptions about the conforma- tional equilibria were made: o\\ “ O <—— 3 ll -’ Ph r C§§ “ r .____9 O h Differentiation between the l- and 4-methyl groupsirs 156 straightforward. The effective shielding constant of Br is larger than COPh, so one would expect the 4-methyl group to appear farther downfield. A difference between 13 axial and equatorial absorption in the nmr and C have been 88 noted previously, where axial protons and methyl groups usually resonate at higher field strengths.157 The conforma- tional assumptions reveal that the l-methyl group is equa- torial most of the time in trans-BDMBC, whereas it is mostly axial for cis-BDMBC. Comparison with similar measurements on 18 and 12 by Lewis158 is revealing. C O\. CH3 61.28 h Q CH3 61.27 h H3 61.43 g r f 13 An analogous effect is noted in the C where the l-methyl group in cis-BDMBC resonates at much higher field strength than trans-BDMBC. In the case of the chloride, the assignment was not so straightforward since only one diastereomer was obtained. However, by analogy with the bromo-derivatives, its physical properties suggested it was the trans-isomer. An X-ray structure (see Figure 13) confirmed the original suspicion. Europium shift studies on the known trans—chloro isomer and the suspected trans-bromo isomer related the structures and 89 confirmed the original assignments (see Figures 14, 15, and 16) . The photoreactivity of each diastereomer also supported the assignments. Identification of Photoproducts All the G-substituted valerophenones produced aceto- phenone and 4-benzoy1-1—butene (4-BB) in varying amounts. In the case of 5-SPh, acetophenone and 4-BB were isolated by preparative vpc and compared to authentic samples. Subse- quent identification of acetophenone and 4-BB was made by comparison of vpc retention times with authentic samples. The radical coupling product from 5-SPh--i.e., diphenyl- disulfide--was identified by comparative vpc retention times. A small peak which was assumed to be the cyclobutanol was observed in the vpc for S-SPh and certain other compounds. The coupling product from 5-SOBu--i.e., BuSOZSBu--was identi- fied by comparative vpc retention times. The olefinic fragments resulting from Type II cleavage were not identified or measured. Elimination of p-methoxyphenol from 5-OPh was not observed. Only acetophenone formation was observed. Cleav- age of a C-O bond in 4-OBzhydl, resulting in the formation of benzhydrol and B-benzoylpropionaldehyde was anticipated but not observed. Again, only Type II cleavage was observed. Unfortunately, BDMBC and CDMBC were found to be extremely unstable to vpc analysis, yielding 1,4-dimethy1-1- benxoylcyclohex-3-ene (DMBC) as a broad, tailing peak upon 90 injection on QF-l, SE-30 or Carbowax columns at various column and injector temperatures. Analogous elimination of HX was observed upon column chromatography on either neutral alumina or silica gel. This decomposition greatly compli- cated the identification of DMBC as a photoproduct. However, prolonged photolysis of CDMBC resulted in complete consumption of starting material and resulted in two peaks in the vpc trace. The shorter retention time peak was shown to be DMBC by comparison with an authentic sample, whereas the longer retention time peak was assumed to be the cyclization product resulting from y-hydrogen abstrac- tion. Attempts to isolate this compound proved futile. However, an IR of the crude photolysate revealed a fairly 1 intense OH absorption at about 3500 cm- . Addition of Br2 to the crude photolysate resulted in the precipitation of an orange solid, which was found to be identical to the bromine 'addition product of DMBC. Furthermore, zinc dust debromina- tion of the orange product yielded DMBC by vpc. Lastly, the photolysis of trans-BDMBC in benzene-d6 containing 0.05M pyridine resulted in the appearance of vinylic proton absorptions in the nmr. Since Type II cleav- age products are not found to any significant extentixtthese 91 systems,159 the vinylic absorptions can be ascribed with a fair degree of confidence to DMBC. Quantum Yields The quantum yields of acetophenone and 4-BB formation from the 6-substituted valerophenones were measured as previously described at 3130A and are presented in Table 16. The solvent effects on II:4-BB ratios are presented in Table 17. Since cis- and trans-BDMBC and trans-CDMBC decomposed upon analysis, yielding DMBC, the product could not be measured directly. Fortunately, trans-CDMBC was found to decompose in consistent amounts to the extent of around 25%. Enhancements beyond this value were taken to be equal to the amount formed upon photolysis. However, cis-enuitrans-BCMBC decomposed quantitatively, so quantum yields were determined by measuring the disappearance of. pyridine, which complexed HBr to form insoluble pyridinium hydrobromide. Benzaldehyde quantum yields from cis- and trans-BDMBC and trans-CDMBC were measured in the usual manner. The quantum yields for forma- tion of DMBC and benzaldehyde from cis-anuitrans-BDMBC and trans-CDMBC are listed in Table 18. Quenching Studies The Stern-Volmer quenching plots for the é-substituted valerophenones were measured as previously described at 31303 with 1,3-pentadiene as quencher and have been listed in Part I. The quenching plot for BDMBC and CDMBC were measured at 36603 by quenching the disappearance of pyridine 92 with napthalene. The formation of benzaldehyde was quenched by napthalene. These values are listed in Tables 18. The effect of temperature upon k t for valerophenone and 6-iodovalerophenone in benzene was determined. The plot is presented in Figure 17. 93 Table 16. Quantum Yields of Acetophenone and 4-Benzoyl-1- Butene Formation from Various 6-Substituted Valero- phenones. Compound ¢II ¢4-BB H (S-X) 15th x, Cl 0.58a 0.10a Br 0.05a 0.55a I . <0.002a 0.43a SCN 0.003 0.25 SAC 0.78 0.02 SBu - 0.21 0.006 SOBu 0.03 0.39 SOZBu 0.39 0.03 StBu 0.16 0.01 SOPh 0.003 0.32 SOZPh 0.19 0.22 SPh 0.015 0.28 0 -.—0Me ---b 0 . 00 ph//\\//\~/’ Ph Y aMeasured by J. H. Sedon in the presence of 0.1M pyridine (see reference 160). b No 4-BB detected. Acetophenone present butrxfi:measured. c:No detectable B-benzoylpropionaldehyde. Acetophenone present but not measured. 94 Table 17. Solvent Effects on B-Cleavage for Various 6-Chloro- valerophenones.a Compound Solvent ¢II ¢4-BB II/4-BB fl Benzene 0.58 0.10 5.8 ./\/\/\Cl Dioxane 0.36 0.06 6.0 CH3CN 0.54 0.08 6.8 MeOH 0.36 0.06 6.0 O Benzene 0.045 0.008 5.6 II MeOH ---b ---b (22) OMe F Benzene 0.63 0.07 9.0 1 Dioxane 0.20 0.028 7.1 CF3 CH3CN 0.63 0.04 16 MeOH 0.43 0.03 14 aA11 measurements made in benzene containing 0.10M pyridine. b Too low to measure. 95 Table 18 . Quantum Yields and Rate Data for 4-Halo-l, 4-Dimethyl- 1-Benzoy1cyclohexanes. Compound a k c"c110 qTCHO ¢DMBC quDMBC Ph 0.008 233 0.21 100 0.01b c O h d 0.019 208 0.45(0.osm) 10.7 0.017 0.40(0.10md Br \Tf::::::;7LIr/Ph 0.11 210 0.12(0.051~4)d 65 B Ph 0.20C 200 0.045 29 E a0.02M 05H. bPresumed to be bicyclic alcohol. cSee reference 161. dNumber in parenthesis refers to the concentration of added pyridine. 96 ago. .mmo_ \ / -J/V ‘ / .mU. .Umzaoimcmuv mo musgosuum Hmummuo >mmix _Go_ .mmoi .eo.//// .moi .HUH -UH _0_ .MH gunmen 97 r, 1 0\ Ph / C1 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 5ppm Figure .14 . NMR Spectra of trans-CDMBC in CDC13 (Bottom) ; in CDCl3 Containing 0.0259g/0.5ml Eu(fod)3(Top). 98 8.0 7.0 6.0 5.0 4.0 3.0 2.0 1.0 GPPm Figure 15 . NMR Spectra of trans-BDMBC in CHCl3 (Bottom) ; in CDCl3 Containing 0.0259g/0.5ml Eu(fod)3(Top). 99 ————— f. L @0011) “J: 10.1.} 8.0 7.0 6.0 5.0 4.0 3.0 2.0 1.0 Gppm Figure 16. NMR Spectra of cis-BDMBC in CDCl3 (Bottom) 7 in CDCl3 Containing 0.02599/0.5ml Eu(fod)3(Top). 100 60“ 40‘ 10 20 30 40 50 60 70 Figure 17. Temperature Dependence of k r for Valerophenone (A) and 6-Iodova1erophenone (A) in Benzene. 101 Discussion Mechanism of Radical B-Cleavage via Photogenerated Diradicals The following transformation was first observed by Wagner and Sedon.162 In light of the known propensities for photoinduced biradicalsix>undergo rearrangements typical of monoradicals, the mechanism seemed straightforward, O O I hv H HX P -——-—>Ph/\/\¢+ x=CLBnI proceeding by radical B-cleavage of the initially formed 1,4-biradical followed by disproportionation as shown below. H H E? MX—e Vii/V 'X _) PM? + HX However, the rate of elimination of Cl deduced from such a mechanism is much faster than that usually assumed.163 In- asmuch as all biradicals possess some zwitterionic charac- ter, anionic elimination as presented below seemed to be a viable mechanistic possibility. (DH (DH HY%\~/Afirf\% ; pW/+\/”\//\y 102 Table 17 presents the quantum yields of ¢II and ¢4-BB for various 6-chlorova1erophenones in different solvents. Little variation in II:4-BB ratios are observed with solvent, except in the case of p-methoxy-G-chlorovalerophenone in CH3CN and MeOH. The ability of polar solvents to stabilize carbonium ions is well known,164 and as such any zwitterionic character of the 1,4-biradical should be stabilized in a polar solvent. Therefore, if X is being eliminated ionically, stabilization of the zwitterion form of the biradical should result in greater amounts of B-elimination relative to Type EIcleavage. Stabilization or destabilization of the zwitterionic form can also be brought about by ring substitution. X Whereas, p-MeO should stabilize the carbonium ion, p-CF3 should destabilize it. In view of the 100 or 1000 fold effect p-substitutents have upon ground state carbonium ion formation,165 the fact that II:4-BB varies little with p-OMe or CF3 in benzene is enlightening. On the other hand, p-CF3 substitution favors Type II elimination slightly in methanol and CH3CN to the extent of a factor of about two. The results for p-OMe-S-Cl in CH3CN and MeOH are confusing in that just the opposite effect is observed. Here, in a situ- ation where B-cleavage should be favored, Type II cleavage predominates. However, the ability of polar solvents to 103 166 and the effect in invert triplet levels is notorious, the p-OMe-S-Cl case is to lower the quantum yields to such an extent that accurate measurements are impossible. Thus, the discrepancy could be due to analytical difficulties. The extent to which p-substitution influences ks, the rate constant of Type II cleavage, which is normally about 107sec-l,167 is unknown; however, it is safe to say that solvent or substituent effects are at best extremely small, and the ionic mode of elimination is probably only a small contributor for only the worst radical leaving groups. Formation of 4-BB or DMBC from their respective precursors via either intramolecular or intermolecularinter- action of the halide or thiyl groups with the excited benzoyl group can be definitely ruled out. Irradiation of 4-SPh resulted in no allylacetophenone formation. 0 hv O Ph/H\/\/SPh +PH/|k/§ + HSPh Interaction between the sulfur and the benzoyl groupsixtthis case was shown in Part I to be faster than S-SPh and should be even more favorable for photolytic elimination by the above mechanism. Intermolecular interaction in the case of BDMBC was ruled out by the following control experiment. r hv acetophenone \\ \ benzene V? 0.5M pyridine 104 This coupled with the fact that alkyl.halides 3"[C T] , + -SPh PhM coupling 27 7% products ° 100% hv Ph/H\/\/\5Ph Type II products Scheme 8. Photolysis <16 6-Thiophenoxyvalerophenone 1J1 Benzene. 106 Quantum Yields and Other Observations Examination of Table 16 and Scheme 8 reveals that in the case of 5-SPh, C-T quenching by sulfur is the major reaction followed by B-elimination to form 4-BB. Inter- estingly, very small yields of PhSSPh were measured, which at best comprised only about 3% of the total products. In-cage coupling of the radicals would result in the situation shown below, which would probably revert to starting material. hjj:~/A\// 1 -——-+' O 'SPh S h However, out-of—cage coupling would be statisticalanuiresult , in the formation of disulfideanuithe corresponding pinacol. (fiIOH OH Ph Ph 2 Ph/i\~/A\¢9’ '—————-9 ;iF—EE / \ The small quantum yield of disulfide coupled with the near 100% material balance allows one to ignore these small side reactions. However, the fact that they are small means that very rapid in-cage disproportionation must be taking place. OH O .SPh /U\/\ Ph , / ——-—) Ph / + SPh 107 In the acyclic cases, the leaving group is already set up for disproportionation as shown in the previous diagram; but in the cyclic case, and especially for the trans- compounds, the leaving group is on the opposite side of the ring with respect to the ketyl. q F‘ Therefore, it must migrate around inside the cage before disproportionation can occur. In-cage molecular reorgani- zation must indeed be very rapid. Rates of B-Elimination Given the plausible assumption that 6-substitution effects little change in kg,169 comparison of II:4-BB ratios allows extraction of relative rates of B-elimination. Table 19 contains the relative rates of B—cleavage for a variety of groups. Comparison of Cl, Bu, and I reveals the expected order of ease of elimination--i.e., I > Br > C1-— which parallels that observed for ionic eliminations.17o The difference between SR and SPh is roughly a factor of 1000 in favor of SPh and can be rationalized in terms of resonance stabilization by the benzene ring, analogous to 108 Table 19. Rate Constants of B-EliminatiOn of Various Substit- uents in G-Substituted Valerophenones. Compound k8,108sec-la kBrel O H 12th E. Cl 0.17 1 Br 1.1 65 I 21.5 1260 SCN 83.0 490 SAC 0.0026 0.15 SBu 0.0029 0.16 SOBu 1.3 76.0 SOzBu 0.0077 0.46 StBu ' 0.0064 0.41 SOPh 10.6 630 SOZPh 0.12 6.8 SPh 1.4 110 aCalculated assuming ks equals 107sec-1. 109 that in a benzylic radical, that is, are ls .- 1, o:- The enhanced rate of elimination of SPh relative to RS has been demonstrated previously171 by the addition of ethane- thiol to allylphenylsulfide as shown below. Here elimination of SPh is faster than chain-transfer with ethanethiol. The three-fold difference in rates between SBu and S-tBu can be explained by increased hyperconjugation in the latter, analogous to that observed for alkyl radicals172 and the fact that the steric bulk of the t-butyl group lowers the bond dissociation energy. Surprisingly, oxidation of the sulfur to a sulfinyl moiety enhances its rate of elimination relative to the sul- fide. Thus, the order PhSO > SOBu >> SOzBu and SBu confirms Kice's173 suggestion that sulfinyl radicals have the greatest kinetic stability of the three. Sulfinyl radicals have been thought to be rather stable by analogy with the isoelectronic nitroxide and dithiyl radicals.174 110 -+ R—S=O 4——) R—S-O- R N—o- (—-—-9 R N—o' R—S—S- 6——) RS$ Interestingly, S-SCN B-cleaves almost exclusively, whereaijt 5-SAc B-cleavage is a minor product. One might expect these radicals to be fairly stable by analogy to allyl radicals. - sea €——> S=C=N vs C—C=C «h—a C=C—C O H - S—CCH 0 l 3 (——> S=C—-CH 3 The known lack of conjugative stabilization in a-keto radicals175 explains the order SCN >> SAc and further indi- cates the X leaves as a radical since ionic cleavage of -SAc should be favorable. The most surprising result of all is that Cl is elimi- nated faster than SR. Kineticists have assumed a much smaller value (If kB for C1, and as such one would expect the observed order to be reversed.176 However, it is possible that previous studies on monoradicals have indi- cated too low a value of k8 due to reverse additions. In the present system rapid in-cage disproportionation tends to minimize the effects of reversible addition. 111 . . . . . . 177 - Competitive elimination studies by Hall are consistent with the present observations. That is, in the addition of ethanethiol to the allylchloride shown below, .0 an unexpected product was observed. . \N EtSvk/Cl EtS ms + // 1 —'7 + Cl 72% 28% The lesser product can be explained by elimination of -C1 followed by ionic addition of HCl to form the observed product. EEEV/L\¢/Cl'———9 EtS '§§4'Cl- 3FEEV/k\y C1 Relative rate constants for B-cleavage have not been previ- ously reported. In fact, the actual rate constants can be estimated by assuming kS equals 107sec-l.178 This would yield kB's of lossec-l and lossec-1 for SBu and Br, respectively. The corresponding G-phenoxyethers are observed not to undergo B-cleavage, indicating that ks >> kB in these Ph/‘(LAA :1, (11A / . no... 112 cases. Cleavage of the comparatively strong C—O bond is apparently unfavorable. In fact, cleavage in 19, which would result in the formation of a C-O double bond and a relatively stable benzhydryl radical, is not observed, reflecting total kinetic control of the reaction. ANO Ph hV N + o P ——> . h P PhAPh 19 o I + P M11 Ph/\Ph y-Hydrogen Abstraction and Anchimeric Assistance The rate constants of y-hydrogen abstraction for 5-Br and S-I have been found to be enhanced by 230% and 530%, respectively, relative to that expected if only inductive 179 The observed enhancements were 180 effects were operative. attributed to anchimeric assistance by the B-haloatom. Numerous examples of B-halo and B-thiyl assistance are known.181 Recently, Shevlinl8 2 presented evidence that suggested a small amount of anchimeric assistance byaphenyl- sulfinyl group in thetflJ1hydride induced elimination of the B-bromophenylsulfinyl group. Since the triplet benzoyl 183 only a group mimics the behavior of alkoxy radicals, minor effect is noted in the case of S-Br and S-I. The highly exothermic nature of abstraction of bromine atoms by 113 tin radicals is probably responsible for the small effect noted in the latter case. A fairly substantial enhancement in ky.was noted in Part I for 5-SOBu--rough1y a factor of ten. Assistance by the sulphinyl group can be envisioned to occur in two dif- ferent ways in which three-(n:four-centered intermediates can be constructed. On the other hand, 5-SBu or 5-SPh show no such enhancements. ‘9 § or (1-18 The concept of anchimeric assistance requires two clarifying observations: one, the lowering of the transi- tive state enthalpy by bridging and greater conformational restrictions on the structure of the transition state (i.e., a more negative AST) require that the temperature dependence of y-hydrogen abstraction be different than that of a com- pound in which no bridging is occuring; and, two, stereo- electronic requirements lead one to expect significant stereochemical effects on the magnitude of anchimeric assis- tance, analogous to those observed by Skell184 for axial and equatorial bromines in a rigid cyclohexyl system. Figure 17 presents the temperature dependence of the qu for 6-iodovalerophenone relative to valerophenone. The fact that the two slopes are different is consistent with the suggestion that anchimeric assistance is operative. 114 The effect of orientation upon the magnitude of anchimeric assistance is striking. Table 20 presents the kY and km of the various diastereomers of BDMBC and CDMBC. 7 l, which is comparable The kY for t-CDMBC is 1.1 - 10 sec- to the kY in 6-chlorovalerophenone. It is faster bya factor of two in the cyclic case due to the fact that the hydrogens are perfectly set up for abstraction. However, kY for trans- 8sec-1, a facter of about 200 times BDMBC is 1.9 - 10 faster. The inductive effect of Br and C1 are comparable18 so the enhanced rate must be due to assistance by Br. Formation of a bridged species as suggested by Skell186 is possible, but the data do run: permit any comment in this respect. The trans-isomer is, of course, perfectly set up for assistance since the C-H and C-Br bonds are trans- periplanar.187 Conformational Effects Scheme 9 depicts the conformational effects in the photochemistry of trans-CDMBC. Lewisl88 has previously shown that when the benzoyl group is axial, only y-abstraction occurs. Conversely, when it is held equatorially, only Type I cleavage occurs. Thus, analogous arguments can be made for the present system. If the respective rates of 115 Table 20. Rate Constants of y-Hydrogen Abstraction and a-Cleavage for the 4-Halo-1,4-Dimethy1-l- Benzoylcyclohexanes. - 1 7 -1 7 - 7 - COmpound ?,10 s kYIlO s ka,10 s c§ Ph (0) 2'14 1.11 0.017 (7) 5.07 1 Ch h (0) 2'40 21.0 0.046 (v)46.7 r h (a) 2.38 0.92 0.26 5' (y) 7.69 Br h (0) 2'50 0.77 0.50 U (v) 17.2 116 Type II Type I products products o§ h -—> c ‘i (— h Cl Th0 hv Q§ Ph ———) C]. H l h C Scheme 9. Conformational Effects in the Photochemistry of trans-CDMBC. 117 reaction--i.e., kY and ka--are faster than equilibration of the excited state conformers, two distinct triplet lifetimes will be observed (kY and ka can, of course,tx3coincidentally equal, in which case a single lifetime will be observed). Thus, product distribution reflects the ground state equili- brium. Lewis189 has found the Type I quantum yield for 22 in the presence of 0.01M dodecanethiol to be 0.31. 20 Assuming that remote substituents have little effect on 0“. comparison of the quantum yields of benzaldehyde formation for the present system should yield the respective ground state population. Thus, utilizing the quantum yields in Table 18, the following ground state conformational equili- bria can be calculated and are presented in Table 21. Obvisouly, the photochemical results and the corre- sponding determination of ground state equilibria agree nicely with the initial predictions and further illustrate how light can be used as a probe. On the other hand, prior equilibration of the two excited conformers would lead to a single lifetime, in which case the product distribution would depend only upon the relative rates of y-hydrogen abstraction and a-cleavage. This, in turn, would reveal little or no information about ground state equilibria. Complications arising from 118 Table 21. Ground State Equilibria of 4-Halo-1,4-Dimethyl- l-Benzoylcyclohexanes. o§ Ph 0 ——-0 Cl I ¢______ Ph Cl 99% 1% (he Ph , 0 '—"° Br ; : 5 H i Ph r 94% 6% (h: Ph Br 0 Br 5 || *— Ph 61% 39% 119 intermolecular energy transfer from one excited conformer to the other are not likely here in light of the low concen- tration used and the short triplet lifetimes. Indications for Further Research Synthetic Utility. The synthetic utility of this reaction as a means of introducing double bonds in a highly regiospecific manner needs to be investigated. Relative Rates of B-Cleavage. This reaction provides an excellent handle for measuring relative rates of k8 for various groups. However, it presents a serious limitation in that kS is too fast to allow elimination of other less stable radicals--i.e., -OPh and -CH2Ph. Perhaps, if ks could be slowed down, the utility of the reaction could be enhanced. Utilization of the thioester derivatives might accomplish this since formation of a C-S double bond is fairly unfavorable. However, a-cleavage would probably be a Type II ///3 products ‘\\\9 B-cleavage 0 OH Ph/\s x ——) Ph - s/\/\x competing reaction. Stabilization of the ketyl radical moiety as shown below might also serve to decrease ks. 120 B-Substituted Butyrophenones. Analogous B-cleavage should also be observable for B-substituted butyrophenones as shown below. OH )( 0 ll PhM __9 PhN§ 0 These systems have not been previously investigated. A cyclic concerted mechanism as mentioned before should be more favorable for this system relative to the 6-substituted valerophenones. Inductive Effects. An investigation into the effects of stereochemistry upon inductive effects would be interesting. For instance, comparison of the following two isomers would allow detection of any difference in the inductive effect upon y-hydrogen abstraction between the axial and equatorial chlorines. 0%. Ph Cl Qb £01 C1 EXPERIMENTAL Preparation and Purification of Materials Solvents and Additives Benzene: (Mallinckrodt) was purified by stirring over concentrated sulfuric acid for several days. The benzene was then washed with additional amounts of sulfuric acid until it remained clear, followed by several washings with water and one final washing with saturated sodium bicar- bonate. The benzene was dried over magnesium sulfate or sodium sulfate and distilled from P205 through :1 column packed with glass helices. Only the middle 50% was collected. Methanol: (Fisher Scientific or Mallinckrodt) was distilled from magnesium turnings. The middle fraction was collected. Dioxane: (Mallinckrodt) was used as received or distilled through a short Vigreux column. Pyridine: (Mallinckrodt) was distilled from barium oxide and the middle fraction collected. Acetonitrile: (Fisher) was distilled rapidly from potassium permanganate. Sulfuric acid was added to the distillate and the distillate was decanted from the ammonium salts. It was then distilled throughaishort Vigreux column. 121 122 Ethanol: was used as received. Hexane: (Mallinckrodt) was used as received. Pentane: (Mallinckrodt or Drake Brothers) was used as received. Benzenethiol: (Aldrich) was used as received. Internal Standards The standards used in this work were purified by Dr. P. J. Wagner as indicated below. Dodecane: (Aldrich) was purified in the same manner as benzene with distillation under reduced pressure. Tetradecane: (Columbia Organics) was purified in the same manner as dodecane. Hexadecane: (Aldrich) was purified in the same manner as dodecane. Heptadecane: (Aldrich) was purified in the same man- ner as dodecane. Octadecane: (Aldrich) was purified by recrystalliza- tion from ethanol. Nonadecane: (Chemical Samples) was purified in the same manner as octadecane. Quenchers Ngthalene: (Matheson Coleman and Bell) was recrystal- lized several times from ethanol. l-Methylnapthalene: (Aldrich) was used as received. 123 Cis- and Trans-1,3-Pentadiene: (Chemical Samples) was used as received. Cis-l,3-Pentadiene: (Chemical Samples) was used as received and found to be 99.8% pure by vpc. n-Butylsulfide: (Aldrich) was used as received. n-Butylsulfoxide: was prepared by 30% hydrogen peroxide oxidation of n-butylsulfide in acetone. It was puri- fied by recrystallization from hexane. n-Butylsulfone: was prepared by 30% hydrogen peroxide oxidation of n-butylsulfide in glacial acetic acid. It was purified by recrystallization from hexane. Ketones Acetophenone: (Matheson Coleman and Bell) was dis- tilled under reduced pressure by A. E. Puchalski. Valerophenone: was prepared by the Friedel-Crafts acylation of benzene by valeryl chloride. The acid chloride was dripped alowly into a mixture of a ten-fold excess of benzene containing a 5% excess of aluminum chloride at O—SOC and stirred for 3-10 hours. The mixture was then poured onto cracked ice and concentrated HCl and the layers separated. The benzene layer was washed several times with dilute HCl, dried over sodium sulfate and evaporated under aspirator pressure. The crude produce was distilled under reduced pressure and the middle fraction collected. Butyrophenone: (Aldrich) was purified by Dr. M. J. Thomas. 124 Phenacylchloride: (Aldrich) was used as received. Phenacylbromide: (Aldrich) was used as received. B-Bromopropiophenone: was prepared by the Friedel- Crafts acylation of benzene with B-bromopropionyl chloride in carbon disulfide at -5-0°C in an analogous manner as valerophenone: NMR(CDC13) 3.2(m,4H), 7.2(m,3H), 7.8(m,2H). y-Chlorobutyrophenone: was prepared by the Friedel- Crafts acylation of benzene by y-chlorobutyryl chloride (Aldrich) in the same manner as valerophenone: NMR(CDC13) 200(m’2H), 300(t’2H), 3.5(t'2H), 7.2(m,3H), 7.8(m,2H). 6-Chlorovalerophenone: was prepared by J. H. Sedon. e-Chlorohexanophenone: was prepared by W. B. Mueller. Phenacylsulfides: (General Procedure) were prepared by treatment of the appropriate phenacyl chloride or bromide with the sodium salt of the corresponding thiol. The appro- priate alkylthiol or arylthiol was added to an ethanolic NaOH solution and stirred for one hour. The phenacyl halide (one equivalent) was added in one portion, and a mild exo- thermic reaction ensued. The mixture was stirred overnight at room temperature and poured into water. It was then extracted with several portions of ether. The combined ether extracts were washed several times with water, dried over sodium sulfate and evaporated. The crude products were either distilled at reduced pressure or recrystallized from ethanol. The following compounds were prepared in this manner . 125 2-Thiomethylacetophenone (Z-SMe): bp 80°C(0.05mm); IR(neat) 2900, 1655, 1360cm"1 ; NMR(CDC13) 62.05(s,3H). 6.05(s,2H), 7.3(m,3H), 7.8(m,2H); m/e 166(M+). 2-Thiobutylacetophenone (2-SBu): bp 120°C(0.50mm); IR(neat) 2950, 1690, 12750m’1; NMR(CDC13) 60.9(m,3H), 1.4(m,4H), 2.5(t,2H), 3.7(s,2H) 7.3(m,3H) 7.8(m,2H); m/e 208(M+). . 2-Thio-t-butylacetophenone (2-StBu): bp 100°C(0.05mm); IR(neat) 2950, 1695, 1280cm-1; NMR(CDC13) 61.3(s,9H), 3.8(s,2H), 7.3(m,3H), 7.8(m,2H); m/e 208(M+). 2-Thiophenylpropiophenone (s-ZSPh): bp 135°C(0.05mm); IR(neat) 3025, 1695, 1230cm—l; NMR(CDC13) 61.5(d,3H), 4.5(m,1H), 7.2(m,8H), 7.8(m,2H); m/e 242(M+). 2-Thiophenylisobutyrophenone (t-ZSPh): bp 130°C (0.07mm); IR(neat) 3025, 1695, 1230001‘1 ; NMR(CDC13) 61.5(s,6H),7.3(m,8H), 8.2(m,2H); m/e 256(M+). Phenyldesylsulfide (PDS): mp 77°C; NMR(CDC13) 65.7(s,lH),7.2(m,13H), 7.8(m,2H); m/e 304(M+). I 2-Thiophenylacetophenone (2-SPh): mp 49°C; IR(CHC13) 3000, 1685, 1605, 120500?1 ; NMR(CDC13) 64.2(s,2H), 7.2(m,8H), 7.8(m,2H); m/e 228(M+). 2-Thiophenyl-4'-fluoroacetophenone (F-ZSPh): bp 150-55°C(0.25mm); IR(neat) 3000, 1685, 1605, 1205cm'1; NMR(CDC13) 64.05(s,2H), 7.08(m,8H), 7.8(m,2H); m/e 246(M+). 2-Thiophenyl-4'-cloroacetophenone (Cl-ZSPh): mp 65°C; IR(CHC13) 3000, 1680, 1600, 12750m'1; NMR(CDC13) 64.1(s,2H), 7.1(m,8H), 7.7(d,2H); m/e 262(M+). 126 2-Thiophenyl-4'-bromoacetophenone (Br-ZSPh): mp 60°C; IR(CHC13) 3000, 1680, 1595, 1280cm‘l ; NMR(CDC13) 64.1(s,2H), 7.1(m,5H), 7.5(m,4H); m/e 307(M+). 2-Thiophenyl-4'-methylacetophenone (Me-ZSPh): nm>61°C; IR(CHC13) 3000, 1660, 1605, 1275cm'1; NMR(CDC13) 62.3(s,3H), 4.1(s,2H), 7.1(m,8H), 7.6(d,2H); m/e 242(M+). 2-Thiophenyl-4'-methoxyacetophenone (OMe-ZSPh): mp 86°C; IR(CHC13) 3000, 1670, 1600, 1265cm-1; NMR(CDC13) 63.7(s,3H), 4.1(s,2H), 6.8(d,2H), 7.1(m,5H), 7.8(d,2H); m/e 258(M+). 2—Thiophenyl-4'-thiomethylacetophenone (SMe-ZSPh): mp 48°C; IR(CHC13) 3000, 1670, 1590, 1095cm-l; NMR(CDC13) 62.4(s,3H), 4.2(s,2H), 7.1(m,8H), 7.6(d,2H); m/e 274(M+). 2-Thiopheny1-4'-cyanoacetophenone (CN-ZSPh): mp 65°C; IR(CHC13) 3000, 2225, 1700, 1275cm-l; NMR(CDC13) 64.2(s,2H), 7.2(s,5H), 7.7(m,4H); m/e 253(M+). 2-Thiophenyl-4'-phenylacetophenone (0-28Ph): mp 92°C; IR(CHC13) 3000, 1700, 1275cm-l; NMR(CDC13) 64.2(s,2H), 7.0-7.5(m,12H), 7.8(d,2H); m/e 304(M+). 2-Thiophenyl-4'-dimethylaminoacetophenone (NMez-ZSPh): was prepared by the reaction of dimethylamine and 2-thiophenyl- 4'-f1uoroacetophenone. 109 of 2-thiophenyl-4'-fluoroaceto- phenone was dissolved in 50ml of xylene and placed in a pres- sure bomb. The bomb was cooled in an ice-salt bath, and 25ml of anhydrous dimethylamine (Aldrich) was added. The bomb was sealed and heated at 80°C for 24 hours with vigorous stirring. The bomb was cooled before opening, and the 127 mixture was worked up as in the general procedure for the, phenacylsulfides. The crude product was crystallized from ethanol and obtained pure in 85% yield: mp 78°C; IR(CHC13) 3000, 1660, 1600, 1375cm-l; NMR(CDC13) 62.9(s,6H), 4.1(s,2H), 6.5(d,2H), 7.1(m,5H), 7.7(d,2H); m/e 271(M+). l-Benzoyl-l-thiophenylcyclopropane (cy-ZSPh): was prepared by the reaction of sodium thiophenoxidevfiifll2-bromo- 4-chlorobutyrophenone and subsequent base cyclization. 28g of 2-bromo-4-chlorobutyrophenone obtained by the bromination of 4-chlorobutyrophenone was dissolved in 100ml of ethanol. To this was added one equivalent of sodium thiophenoxide, and this was allowed to stir for four hours. A slight excess of sodium methoxide was added in one portion to effect cyclization, and the product worked up as in the phenacyl- sulfides. Recrystallization from ethanol gave a 91% yield of pure product: mp 62°C; NMR(CDC13) 61.3(m,2H), 1.7(m,2H), 7.2(m,8H), 7.8(m,2H); m/e 254(M+). 4'-thiophenylmethylacetophenone (4'-MSPh): was pre- pared from sodium thiophenoxide and 4'-bromomethylacetophe- none. 4'-bromomethylacetophenone was obtained by irradiation of a mixture of 40g 4'-methylacetophenone and 53g n-bromo- succinimide in 500ml carbon tetrachloride. Pure bromide was obtained by fractionation at reduced pressure in about 30% yield. The reaction of sodium thiophenoxide and 4'-bromo- acetophenone was carried out in the same manner as the phenacylsulfides. Pure product was obtained from ethanol in a 95% yield: mp 91°C; IR(CHC13) 3000, 1685, lzsscm‘l; 128 NMR(CDC13) 62.5(s,3H),-4.l(s,2H), 7.2(m,8H), 7.7(d,2H); m/e 242(M+). Benzoylsulfides: (General Procedure) were prepared essentially the same as the phenacylsulfides except that the reaction mixtures were refluxed for a period of 5-12 hours. In the case of 6-benzoylsulfides, the carbonyl group of 4-chlorobutyrophenone was protected by forming the ethylene glycol ketal prior to reaction with sodium thiolates; other- wise, good yields of phenylcyclopropylketone were obtained. Intramolecular cyclization of the haloketones was only found in the case of 4-chlorobutyrophenone. The ketal was then hydrolyzed by vigorous stirring in 15% HCl overnight. The products were either distilled under reduced pressure or recrystallized from ethanol. The following compounds were prepared in this manner. 3-Thiobutylpropiophenone (3-SBu): bp l30-150°C (aspirator); IR(neat) 2950, 1680, 1450, 135000‘1; NMR(CDC13) 60.9(m,3H), l.4(m,4H), 2.4-3.4(m,6H), 7.3(m,3H), 7.8(m,2H); m/e 222(M+). 4-Thiobutylbutyrophenone (4-SBu): bp 145°C(0.3mm); IR(neat) 2950, 1695, 1450, 12300m‘1; NMR(CDC13) 60.9(m,3H), l.4(m,4H), 1.9(m,2H), 2.5(m,4H), 3.l(t,2H), 7.3(m,3H), 7.8(m,2H); m/e 236(M+). 4-Thio-t-butylbutyrophenone (4—StBu): bp 125°C (0.15mm); IR(neat) 2950, 1695, 1225cm’1: NMR(CDC13) 61.25(s,9H), 2.0(m,2H), 2.6(t,2H), 3.1(t,2H), 7.3(m,3H), 7.8(m,2H); m/e 236(M+). 129 4-Thiophenylbutyrophenone (4-SPh): mp 35°C; IR(CHC13) 2990, 1675, lzzscm"l ; NMR(CDC13) 62.0(m,2H), 3.0(M,4H), 7.2(m,8H), 7.8(m,2H); m/e 254(M+). 5-Thiobutylvalerophenone (5-SBu): bp 155°C(0.3mm); IR(neat) 2930, 1685, 1450, lzzocm"l ; NMR(CDC13) 60.9(m,3H). l.6(m,8H), 2.4-3.0(m,6H), 7.3(m,3H), 7.8(m,2H); m/e 250(M+). S-Thio-t-butylvalerophenone (S-StBu): bp 140°C (0.06mm); IR(neat) 2950, 1695, 1220cm'1; NMR(CDC13) 61.2(s,9H), l.7(m,4H), 2.5(t-2H), 2.9(t-2H), 7.3(m-3H), 7.8(m—2H): m/e 250(M+). 5-Thiopheny1valerophenone (5-SPh): mp 87°C; IR(CHC13) 2995, 1675, 1200cm'l ; NMR(CDC13) 51.9(m,4H) , 2.9(t,4H) , 7.3(m,8H), 7.8(m,2H); m/e 270(M+). 6-Thiobutylhexanophenone (6-SBu): bp 145-160°C (0.4-0.5mm); IR(neat) 2930, 1695, 1450, 1220001“1 ; NMR(CDC13) 6009_, 1781 (1970). H. L. J. Backstrom and K. Sandros, Acta. Chem. Scand., 22, 958 (1962). P. J. Wagner and I. Kochevar, J. Am. Chem. Soc., _2, 2232 (1968). P. J. Wagner, J. M. McGrath, and R. G. ZePP; ibid., 22, 6883 (1972). 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 149 D. O. Cowan and A. A. Baum, ibid., 22, 1153 (1971). P. J. Wagner, J. M. McGrath, and R. G. Zeppr ibid., 24, 6883 (1972). _" J. D. Coyle, Chem. Rev., 12, 97 (1978). P. J. Wagner, Top. Curr. Chem., 22, 1 (1976). F. D. Lewis and T. A. Hillard, J. Am. Chem. Soc., 22, 3852 (1972). P. J. Wagner and R. G. Zepp, ibid., 22, 287 (1972). J. A. Barltrop and J. D. Coyle, "Excited States in Organic Chemistry," John Wiley and Sons, Inc., New York, 1975, p. 180. P. J. Wagner and J. McGrath, J. Am. Chem. Soc., 22, 3849 (1972). H. G. Heine, Justus Liebigs Ann. Chem., 732, 165 (1970). F. D. Lewis and J. G. Magyar, J. Org. Chem., 22, 3102 (1972). F. D. Lewis and R. W. Johnson, J. Am. Chem. Soc., 22, 8915 (1972). J. D. Coyle, Chem. Rev., 76, 97 (1978). E. S. Huyser, "Free-Radical Chain Reactions," John Wiley and Sons, Inc., New York, 1970, p. 211. H. G. Heine, J. J. Rosenkronz, and H. Rudolph, Angew. Chem. Intern. Ed. Eng., 22, 974 (1972). G. A. Berchtold and W. C. Lumma, J. Org, Chem., 22, 1566 (1969). J. R. Collier and J. Hill, Chem. Commun., 640 (1969). S. Majeti, Tetrahedron Letters, 2323 (1971). P. deMayo, C. L. McIntosh, and R. W. Yip, Tetrahedron Letters, 37 (1967). P. J. Wagner, "Creation and Detection of the Excited State," Vol. 1A, A. A. Lamola, ed., Marcel Decker, New York, 1971, p. 173. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 150 A. A. Lamola and G. S. Hammond, J. Chem. Phys., 22, 2129 (1965). P. J. Wagner, 1. E. Kochevar, and A. E. Kemppainen, J. Am. Chem. Soc., 22, 7489 (1972). A. A. Lamola and G. S. Hammond, J. Chem. Phys., 22, 2129 (1965). P. J. Wagner, Accounts Chem. Res., 2, 168 (1971). P. J. Wagner and R. G. Zepp, J. Am. Chem. Soc., 22, 4958 (1971). P. J. Wagner, P. A. Kelso, A. E. Kemppainen, and R. G. Zeppr ibid., 22, 7500 (1972). S. an (editor), "Organic Chemistry of Sulfur," Plenum Press, New York, 1977, pp. 249, 393, and 528. P. J. Wagner, Tetrahedron Letters, 1753 (1967). P. J. Wagner, P. A. Kelso, A. E. Kemppainen, J. M. McGrath, H. N. Schott, and R. G. Zepp, J. Am. Chem. Soc., 22, 7506 (1972). A. Padwa and D. Pashayan, J. Org. Chem., 22, 3550 (1971). F. D. Lewis and N. J. Turro, J. Am. Chem. Soc., 22, 311 (1970). M. C. Caserio, W. Lauer, and T. Novinson, ibid., 22, 6082 (1970). P. J. Wagner and A. E. Kemppainen, ibid., 22, 7495 (1972). C. Walling and M. J. Mintz, ibid., 89, 1515 (1967). A. Ohno, Y. Ohnishi, and N. Kito, Int. J. Sulfur Chem., 2, 153 (1971). E. S. Huyser, "Free-Radical Chain Reactions," John Wiley and Sons, New York, 1970, p. 235. J. K. Kochi and P. J. Krusic, J. Am. Chem. Soc., 22, 3940 (1969). P. J. Wagner and A. E. Kemppainen, ibid., 22, 7495 (1972). -—-—- F. D. Lewis andN.J.Turro, ibid., a, 311 (1970). 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 151 P. J. Wagner and A. E. Kemppainen, ibid., 22, 7495 (1972). P. J. Wagner and A. E. Kemppainen, ibid., 22, 7495 (1972). P. J. Wagner and J. H. Sedon, Tetrahedron Letters, 1927 (1978). E. S. Huyser and R. H. C. Feng, J. Org. Chem., _2, 731 (1971). P. B. Shevlin, T. E. Boothe, J. L. Greene, M. R. Willcoth, R. R. Inners, and A. Cornelis, J. Am. Chem. Soc., 100, 3874 (1978). (a) J. R. Collier and J. Hill, Chem. Commun., 700 (1968). (b) P. deMayo, C. L. McIntosh, and R. W. Yip, Tetra- hedron Letters, 37 (1967). F. D. Lewis, private communication. P. J. Wagner and J. M. McGrath, J. Am. Chem. Soc., 22, 3849 (1972). F. D. Lewis, C. H. Hoyle, and J. G. Magyar, J. Org. Chem., 22, 488 (1975). F. D. Lewis, R. T. Lanterback, H. G. Heine, W. Hart- mann, and H. Rudolph, J. Am. Chem. Soc., 22, 1519 (1975). S. Majeti, Tetrahedron Letters, 2523 (1971). P. deMayo, C. L. McIntosh, and R. W. Yip, ibid., 22, (1967). M. A. El-Sayed, Accounts Chem. Res., 2, 8 (1968). B. J. Scheve, Ph.D. Thesis, Michigan State University, 1974. P. J. Wagner and A. E. Kemppainen, J. Am. Chem. Soc., 22, 7495 (1972). P. J. Wagner and A. E. Kemppainen, ibid., 22, 7495 (1972). t W. B. Mueller, unpublished results. P. J. Wagner and B. J. Scheve, J. Am. Chem. Soc., 22, 1858 (1977). 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 152 P. J. Wagner and A. E. Kemppainen, ibid., 22, 3085 (1969). P. J. Wagner and D. A. Ersfeld, ibid., 22, 4515 (1976). A. Padwa and D. Pashayan, J. Org. Chem., 22, 3550 (1971). A. H. Weller and H. Leonhard, Ber. Bunsenges. Phys. Chem., 21, 791 (1963). S. G. Cohen and J. B. Guttenplan, Chem. Commun., 247, (1969). S. G. Cohen and J. B. Guttenplan, J. Org. Chem., 22, 2001 (1973). C. R. Johnson, H. Diefenbach, J. E. Kaiser, and J. C. Sharp, Tetrahedron, 22, 5649 (1969). N. Kharasch (editor), "Organic Sulfur Compounds," Vol I, Pergammon Press, New York, 1961, pp. 47-74. A. J. Gordon and R. A. Ford, "Chemist's Companion," Wiley-Interscience, New York, 1972, p. 238. H. H. Jaffe and M. Orchin, "Theory and Application of Ultraviolet Spectroscopy," John Wiley and Sons, New York, 1962. S. G. Cohen and J. B. Guttenplan, Chem. Commun., 247 (1969). S. G. Cohen and J. B. Guttenplan, J. Org. Chem., 22, 2001 (1973). (a) E. L. Eliel, "Steric Effects in Organic Chemistry," M. S. Newman, ed., John Wiley and Sons, New York, 1956, pp. 114-120. (b) E. L. Eliel, "Stereochemistry of Carbon Compounds," McGraw Hill, New York, 1962, pp. 198-202. J. E. Gano and L. Eizenberg, J. Am. Chem. Soc., 22, 974 (1973). F. D. Lewis and T. A. Hilliard, J. Am. Chem. Soc., 22, 6672 (1970). P. J. Wagner and M. J. Thomas, ibid., 8, 241 (1976). G. S. Hammond and P. A. Leermakers, ibid., 22, 207 (1962). 153 110. P. J. Wagner and D. A. Ersfeld, ibid., 22, 4515 (1976). 111. S. Majeti, Tetrahedron Letters, 2523 (1971). 112. P. J. Wagner, A. E. Kemppainen, and H. N. Schott, J. Am. Chem. Soc., 22, 5604 (1973). .113. J. A. Barltrop and A. J. Thomson, J. Chem. Soc. (C), 155 (1968). 114. K. Schaffner and S. Iwasaki, Helv. Chim, Acta., 22, 557 (1968). 115. P. S. Skell and P. D. Readio, J. Am. Chem. Soc., 22, 3334 (1964). 116. G. Karabatsos in "Topics in Stereochemistry," E. L. Eliel and N. L. Allinger, eds., Vol V, 1970, pp. 167- 203. 117. G. Karabatsos and O. J. Fenoglio, J. Am. Chem. Soc., 22, 3577 (1969). 118. N. L. Allinger, J. Allinger, L. A. Freiberg, R. F. Czaja, and N. A. LeBei, J. Am. Chem. Soc., 22, 5876 (1960). 119. K. Schaffner and S. Iwasaki, Helv. Chim, Acta.,22, 557 (1968). 120. J. Bitha and J. Hlavka, Tetrahedron Letters, 3843 (1966). 121. J. Schriesheim, Chem. Ind., (London), 1234 (1963). 122. P. G. Russel, J. Phys. Chem., 22, 1350 (1975). 123. S. Majeti, Tetrahedron Letters, 2523 (1971). 124. P. J. Carlson and K. U. Ingold, J. Am. Chem. Soc., 22, 1055 (1968). 125. S. Huyser and J. D. Taliaferro, J. Org. Chem., NM 8, 3442 (1963). 126. R. A. Sheldon and J. K. Kochi, J. Am. Chem. Soc., 22, 4392 (1970). 127. J. Calvert and J. Kraus, ibid., 12, 5921 (1957). 128. P. D. Bartlett and S. F. Nelson, ibid., 22, 135 (1966). 154 129. D. R. Kearns and W. A. Case, ibid., 22, 5087 (1966). 130. P. J. Wagner, A. E. Kemppainen, and H. N. Schott, ibid., 22, 5604 (1973). 131. P. J. Wagner, A. E. Kemppainen, and H. N. Schott, ibid., 22, 5604 (1973). 132. Y. H. Li and E. C. Lim, Chem. Phys, Lett., 2, 15 (1970). 133. P. J. Wagner,bL.May, and A. Haug, ibid., 22, 545 (1972). ‘_—" 134. P. J. Wagner, M. May, and A. Haug, ibid., 22, 545 (1972). 135. R. Castonguay and A. C. Albrecht, ibid., 2, 89 (1970). 136. E. J. Corey, J. Am. Chem. Soc., 75, 2301 (1953). 137. R. N. Jones, D. A. Ramsey, F. Herling, and K. Dobriner, ibid., 12, 2828 (1952). 138. N. L. Allinger, J. C. Tai, and M. A. Miller, ibid., 22, 4495 (1966). 139. R. Hoffmann, C. C. Levin, W. J. Hehre, and J. Hudec, J. Chem. Soc., Perkin II, 210 (1973). 140. J. N. Murrell, "The Theory of the Electronic Spectra of Organic Molecules," John Wiley and Sons, New York, 1963, pp. 164-168. 141. J. Labhart and G. Wagniere, Helv. Chim. Acta, 22, 2219 (1959). 142. R. F. Whitlock and A. B. F. Duncan, J. Chem. Phys., 22, 218 (1971). 143. R. C. Cookson, J. Henstock, J. Hudec. J. Am. Chem. Soc., 22, 1061 (1966). 144. W. D. Chandler and L. Goodman, J. Mol. §pec.,22, 232 (1970). 145. P. J. Wagner, A. E. Kemppainen, and H. N. Schott, J. Am. Chem. Soc., 22, 5604 (1973). 146. P. J. Wagner and A. E. Kemppainen, ibid., 22, 5898 (1968). 147. 148. 149. 150. 151. 152} 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163. 155 P. J. Wagner, Accounts Chem. Res., 2, 168 (1971). G. A. Berchtold and W. C. Lumma, 1566 (1969). J. Org. Chem., 22, C. Walling, and Sons, New York, 1957, pp. "Free Radicals in Solution," John Wiley 467-563. C. Walling and P. S. Fredricks, J. Am. Chem. Soc., 22, 3326 (1962). M. Szwarc and D. Williams, J. Chem. 1171 (1952). Phys., 39. K. Griesbaum, Angew. Chem. Int. Ed., 2, 273 (1970). J. L. Rice in "Free Radicals," J. K. Kochi, ed., Wiley Interscience, New York, Vol. II, p. 809. s. w. Benson, Chem. Rev., lg, 23 (1978). E. L. Eliel, McGraw-Hill, "Stereochemistry of Carbon Compounds," Inc., 1962, p. 236. D. J. Pasto and C. R. Johnson, Determination," Prentice-Hall, Cliffs, 1969, p. 169. "Organic Structure Inc., Englewood R. M. Silverstein, G. C. Bassler, and T. C. Morrill, "Spectrometric Identification of Organic Compounds," John Wiley and Sons, Inc., New York, 1974, p. 168. F. D. Chem. Lewis, R. W. Johnson, and D. E. Johnson,.J.Am. Soc., 22, 6090 (1974). F. D. ibid., 94, Lewis and R. W. Johnson, 8915(1972). P. J. Wagner and J. H. Sedon, Tetrahedron Letters, 1927 (1978). F. D. Lewis, R. W. Johnson, and D. E. Johnson,.J.Am. Chem. Soc., 22, 6090 (1974). P. J. Wagner and J. H. Sedon, Tetrahedron Letters, 1927 (1978). (a) R. Eckling, P. Goldfinger, G. Huybrechts, G. Martens, L. Meyers, and S. Smoes, Chem. Ber., 22, 3014 (1960). (b) P. B. Ayscough, A. J. Cocker, F. S. Dainton, S. Hirst, and M. Weston, Proc. Chem. Soc. London, 244 (1961). 164. 165. 166. 167. 168. 169. 170. 171. 172. 173. 174. 175. 176. 177. 178. 179. 180. 156 N. S. Isaacs, “Reactive Intermediates in Organic Chemistry," John Wiley and Sons, New York, 1974, p. 105. H. C. Brown and Y. Okamoto, J. Am. Chem. Soc., 22, 1913 (1957). (a) A. A. Lamola, J. Chem. Phys., _4_7_, 4810 (1967). (b) R. D. Rauh and P. A. Leermakers, J. Am. Chem. Soc., 22, 2246 (1968). R. D. Small and J. C. Scarano, Chem. Phys. Lett., 22, 431 (1977). P. J. Wagner, J. H. Sedon, and M. J. Lindstrom, J. Am. Chem. Soc., 100, 2579 (1978). H. E. O'Neal, R. G. Miller, and E. Gunderson, ibid., '2_, 3351 (1974). H. C. Brown and G. Wheeler, ibid., 12, 2199 (1956). K. Griesbaum, Angew. Chem. Int. Ed., 2, 273 (1970). E. S. Huyser, "Free-Radical Chain Reactions," Wiley- Interscience, New York, 1970, p. 229. J. L. Kice in "Free Radicals," J. K. Kochi, ed., Wiley-Interscience, New York, Vol. II, p. 715. E. Block, Quart. Reports on Sulfur Chem., 2, 315 (1969). G. A. Russell and J. Lokensgard, J. Am. Chem. Soc., 22, 5059 (1967). P. B. Ayscough, A. J. Cocker, F. S. Dainton, S. Hirst, and M. Weston, Proc. Chem. Soc. London, 244 (1961). D. N. Hall, J. Org. Chem., 2, 2082 (1967). R. D. Small and J. C. Scaiano, Chem. Phys. Lett., 22, 431 (1977). P. J. Wagner and J. H. Sedon, Tetrahedron Letters, 1927 (1978). P. J. Wagner and J. H. Sedon, ibid., 1927 (1978). 181. 182. 183. 184. 185. 186. 187. 188. 189. 190. 191. 192. 193. 157 P. S. Skell and K. J. Shea in "Free Radicals," J. K. Kochi, ed., Wiley-Interscience, New York, Vol. II, p. 809. P. B. Shevlin, T. E. Booth, J. L. Greene, M. R. Willcott, R. R. Inners, and A. Cornelis, J. Am. Chem. Soc., 100, 3874 (1978). C. Walling and M. J. Gibian, J. Am. Chem. Soc., 21, 3361 (1965). P. S. Skell and P. D. Readio, ibid., 86, 3334 (1964). L. P. Hammet, "Physical Organic Chemistry," McGraw- Hill, Inc., New York, 1970, p. 384. P. S. Skell, P. D. Readio, R. Tuleen, J. Am. Chem. Soc., 22, 2615 (1966). P. S. Skell and P. D. Readio, ibid., 86, 3334 (1964). F. D. Lewis and R. S. Johnson, ibid., 22, 8915 (1972). F. D. Lewis, R. W. Johnson, and D. E. Johnson, ibid., 96, 6090 (1974). E. J. Corey and M. Chaykovsky, ibid., 22, 1345 (1965). B. J. Scheve, Ph.D. Thesis, Michigan State University, 1974. S. Majeti, Tetrahedron Letters, 2523 (1971). P. J. Wagner, I. E. Kochevar, and A. E. Kemppainen, J. Am. Chem. Soc., 22, 7489 (1972). APPENDIX APPENDIX This section contains the raw experimental data, such as product to standard ratios, internal standard concentrations, etc., used to calculate the various photokinetic parameters. The data is arranged in tabular form as follows: the concentration of internal standard for the actinometry is listed first. C16 was used exclusively for all valero- phenone actinometry. The correction factor for acetophenone versus C16 is 2.3. Other correction factors are listed (in parentheses) after the concentration of standard to which it refers. The quencher that was used for the Stern-Volmer plot is listed next, followed by the column and column conditions. The description of the various columns are listed in the Experimental section. Quenching data, including quencher concentrations, product to standard ratios, and ¢°/¢ values, are listed next. The concentrations of internal standards for the quenching 3 to lo'ZM. runs were not measured but were in the range 10- The values immediately below the quenching data are the prod/std ratios used for calculating the respective quantum yields. In most cases the concentration of standard in all tubes is the same as the actinometer. In some cases 158 159 different concentrations of standard were used for each tube. In these cases the prod/std ratios are listed in the same order as previously listed concentrations. Abbreviations used in this section are: Pip = piperylene; l-NM = l-methylnapthalene; Napth = napthalene; act = actinometer (valerophenone in all cases); acet = acetophenone; Bzald = benzaldehyde; DMBC = 1,4-dimethyl-l- benzoylcyclohexane; 4-BB = 4-benzoyl-l-butene; and pyr = pyridine. All quantum yields were performed at 3130A. Quenching runs utilizing piperylene were performed at 31303. Quenching runs utilizing napthalene or l-methyl- napthalene were performed at 3660A. Photolysis times were usually between one and three hours. The intensity of the lamp varied between 0.005 and 0.01 Einsteins M71. 160 Table 22. Data for 2-Thiomethylacetophenone. 0.0140M C16, 0.0154M C16, 0.0139M C16 Pip quencher Col #1 145°C [Q] ' prod/std ¢°/¢ 0.00 2.38 -- 0.24 1.40 1.70 0.55 0.96 2.30 1.26 0.66 3.60 1.53 0.59 4.10 act 0.158 acet 0.122 acet(0.05M PhSH) 0.169 Table 23. Data for Thio-t-butylacetophenone. 0.0139M C16 l-MN quencher Col #1 145°C [0] prod/std ¢°/¢ 0.00 0.88 -- 0.08 0.68 1.29 0.16 0.53 1.66 0.50 0.29 3.00 act 0.140 acet(0.05M PhSH) 0.017 161 Table 24. Data for 2-Thiobutylacetophenone. 0.025M C16 l-MN quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.0 0.49 -- 0.3 0.29 1.67 0.7 0.24 2.00 1.0 0.19 2.50 act 1.82 acet 2.37 acet(0.05M PhSH) 2.84 Table 25. Data for 4-Thiobutylbutyrophenone. 0.011M C16 Pip quencher Col #1 145°C acet(l.0M dioxane) [Q] prod/std ¢°/¢ 0.0 0.20 -- 0.3 0.13 1.5 0.6 0.10 1.9 1.0 0.08 2.6 act 0.27 acet 0.08 0.15 162 Table 26. Data for 4-Thio-t—butylbutyrophenone. 0.014M C16 Pip quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.0 0.29 -- 0.4 0.16 1.8 0.8 0.09 2.9 1.2 0.07 3.6 act 0.49 acet 0.39 Table 27. Data for 4-Thiophenylbutyrophenone. 0.005M C16 Pip quencher Col #1 145°C acet(l.0M dioxane) [Q] prod/std ¢°/¢ 0.0 0.37 -- 0.3 0.16 2.4 0.6 0.09 3.9 1.0 0.06 5.9 act 0.52 acet 0.50 0.57 163 Table 28. Data for 5—Thiobutylvalerophenone. 0.006M C12 (1.63) Pip quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.00 0.72 -- 0.02 0.44 1.6 0.06 0.22 3.3 0.10 0.16 4.5 act 0.330 acet(l.0M dioxane) 0.210 4-BB(l.0M dioxane) 0.003 Table 29. Data for 5-Thiophenylvalerophenone. 0.011M C16 (1.67) Pip quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.00 0.72 -- 0.02 0.44 1.6 0.06 0.22 3.3 0.10 0.16 4.5 act 0.615 acet(l.0M dioxane) 0.029 4-BB(l.0M dioxane) 0.840 .— 164 Table 30. Data for 6-Thicbuty1hexanophenone. 0.026M C16 Pip quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.00 0.32 -- 0.02 0.20 1.6 0.06 0.09 3.3 0.10 0.06 5.2 act 0.16 acet 0.10 acet(1.0M dioxane) 0.12 Table 31. Data for 5-Thiocyanatovalerophenone. 0.041M C16 (1.67) Pip quencher C01 #1 145°C [Q] prod/std ¢°/¢ 0.0 2.0 -- 0.2 0.64 3.10 0.4 0.39 5.10 0.6 0.17 11.0 act 1.38 acet(l.0M dioxane) 0.02 4-BB(1.0M dioxane) 1.27 165 Table 32. Data for 5-Thioacetoxyva1erophencne. 0.0548M C16 (1.67) Pip quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.00 1.61 -- 0.06 0.42 3.8 0.12 0.26 6.0 act 0.370 acet 0.023 acet(1.0M pyridene) 0.870 4-BB 0.648 Table 33. Data for 2-Methy1sulfinylacetophenone. 0.025M C16 1-MN quencher Col #1 145°C acet(0.05M PhSH) [Q] prod/std ¢°/¢ 0.0 1.58 -- 0.5 1.06 1.48 1.0 0.80 1.97 act 1.13 acet 0.37 1.49 166 Table 34. Data for 4-Buty1sulfinylbutyrophenone. 0.011M C16 Pip quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.00 0.52 -- 0.02 0.41 1.28 0.04 0.29 1.78 0.06 0.27 1.89 0.08 0.23 2.27 act 0.516 acet 0.048 acet(1.0M dioxane) 0.052 Table 35. Data for 5-Buty1sulfinylvalerophenone. 0.0088M C16 (1.67) Pip quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.00 0.78 -- 0.02 0.52 1.5 0.04 0.41 1.9 0.06 0.35 2.3 0.08 0.29 2.7 act ' 0.56 acet(1.0M dioxane) 0.05 4—BB(1.0M dioxane) 0.92 167 Table 36. Data for 2-Butylsu1fcnylacetophenone. 0.021M C16 l-MN quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.000 0.63 -- 0.012 0.19 3.2 0.016 0.14 4.6 0.020 - 0.11 5.7 act 1.78 acet 1.05 acet(0.05M PhSH) 0.28 Table 37. Data for S-Butylsulfonylbutyrophenone. 0.0075M C16 Pip quencher C01 #1 145°C [Q] prod/std ¢°/¢ 0.0000 1.75 -- 0.0004 0.746 2.3 0.0008 0.481 3.6 0.0010 0.300 5.8 act 0.932 acet 0.566 acet(1.0M dioxane) 0.263 168 Table 38. Data for 5-Butylsu1fony1va1erophenone. 0.0085M C16 (1.67) Pip quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.000 1.36 -- 0.002 1.10 1.24 0.004 0.75 1.81 0.006 0.61 2.23 act 0.66 acet(l.0M dioxane) 0.78 4-BB(1.0M dioxane) 0.07 Table 39. Data for 2-Thiopheny1acetophenone. 0.016M C16 1-MN quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.0 0.53 -- 2.0 0.51 1.0 act 0.220 acet 0.051 acet(0.05M PhSH) 0.162 169 Table 40. Data for 2-Thiophenyl-4'-F1uoroacetophenone. 0.012M C16(2.3) 1-MN quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.0 0.6 -- 2.0 0.6 1.0 act 0.45 acet . 0.063 acet(0.05M PhSH) 0.312 Table 41. Data for Thiophenyl-4'-ch10rcacetophenone. 0.012M C14, 01014M C14 (2.0) Napth quencher Col #1 145°C acet(0.05M PhSH) [Q] prod/std ¢°/¢ 0.00 0.52 -- 0.44 0.38 1.36 0.86 0.32 1.61 1.76 0.21 2.41 act 0.450 acet 0.075 0.344 170 Table 42. Data for 2-Thiopheny1-4'-bromoacetophenone. 0.012M C16, 0.0136M C16 (2.0) Napth quencher Col #1 145°C [Q] prod/std ¢°/¢' 0.00 1.01 -- 0.56 0.51 1.78 0.96 0.39 2.61 1.45 0.27 3.70 act 0.450 acet 0.085 acet(0.05M PhSH) 0.265 Table 43. Data for 2-Thiopheny1-4'-methy1acetophenone. 0.012M C16, 0.013M C14 (1.8) Napth quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.0 0.32 -- 2.0 0.31 1.0 act 0.45 acet 0.078 acet(0.05M PhSH) 0.462 171 Table 44. Data for 2-Thiopheny1-4'-methoxyacetophenone. 0.0026M C16 (2.04) Napth quencher Col #1 145°C [Q] prod/std ¢°/¢ \ 0.00 0.39 -- 0.48 0.20 1.97 0.98 0.12 3.36 act 0.263 acet 0.078 acet(0.05M PhSH) 0.367 Table 45. Data for 2-Thiopheny1-4 '-thiomethoxyacetophenone. 0.011M C16, 0.007M C19 (2.09) Napth quencher Col #1 170°C [Q] prod/std ¢°/¢ 0.000 0.57 -- 0.007 0.38 1.48 0.040 0.24 2.32 0.090 0.15 3.68 act 0.39 acet 0.09 acet(0.05M PhSH) 0.42 172 Table 46. Data for 2-Thiopheny1-4'-dimethylaminoacetophenone. 0.011M C16, 0.002M C20 (2.37) Napth quencher C01 #1 190°C [Q] prod/std ¢°/¢ 0.00 1.68 -- 0.04 1.21 1.34 0.06 0.900 1.87 0.10 0.600 2.80 act 0.210 acet 0.267 acet(0.05M PhSH) 0.443 Table 47. Data for 2-Thiopheny1-4'-pheny1acetophenone. 0.016M C16, 0.005M C20 (2.3) Napth quencher Col #1 190°C acet(0.05M PhSH) [Q] Prod/std ¢°/¢ 0.000 4.43 -- 0.008 2.99 1.48 0.015 2.57 1.72 0.026 2.17 2.00 act 0.630 acet 0.270 1.670 173 Table 48. Data for 2-Thiopheny1-4'-cyanoacetophenone. 0.0126M C16,0.0020M C19 (1.72) Napth quencher Col #1 190°C [Q] prod/std ¢°/¢ 0.00 0.44 -- 0.16 0.34 1.29 0.56 0.29 1.52 0.96 0.22 2.00 act 0.31 acet 0.25 acet(0.05M PhSH) 1.14 Table 49. Data for 2-Thiophenylpropriophenone. 0.027M C16 (2.04) Napth quengher C01 #1 145 C [Q] prod/std ¢°/¢ 0.0 0.61 -- 2.0 0.60 1.0 act 0.580 acet 0.256 acet(0.05M PhSH) 0.637 174 Table 50. Data for 2-Thiophenylisobutyrophenone. 0.0026M C16 (1.84) Napth quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.0 0.24 -- 2.0 0.25 1.0 act 0.49 acet 0.67 acet(0.05M PhSH) 0.69 Table 51. Data for 1-Thiophenyl-1-benzoy1cyclopr0pane. 0.013M C16, 0.015M C15 (1.89), 0.014M C15(1d89) Pip quencher Col #1 145°C [Q] ' prod/std ¢°/¢ 0.00 0.77 -- 1.12 0.33 2.33 1.65 0.20 3.61 act 0.274 acet 0.119 acet(0.05M PhSH) 0.382 175 Table 52. Data for 4'-Thiopheny1methylacetophenone. 0.0136M C16, 0.0070M C18 (2.3), 0.0063MC18 (2.3) Napth quencher Col #1 145°C [Q] prod/std ¢°/¢ 0.0 0.19 -- 2.0 0.17 1.0 act 0.29 acet 0.19 acet(0.05M PhSH) 0.77 Table 53. Data for 4'-Thiopheny1acetophenone. 0.012SM C16, 0.0137M C16 C01 #1 145°C prod/std act 0.64 acet(0.05M PhSH) 0.001 176 Table 54. Data for Phenyldesylsulfide. 0.0126M C16, 0.0090MC20 (1.5) , 0.0100MC20 (1.5) Col #1 190°C prod/std act 0.14 acet 0.0018 acet(0.05M PhSH) 0.0144 Table 55. Data for 5-Pheny1sulfinylvalerophenone. 0.0028M C16 (1.67) C01 #1 145°C prod/std act 1.66 acet 0.009 acet(1.0M dioxane) 0.010 4-BB 1.770 4-BB(1.0M dioxane) 1.730 177 Table 56. Data for S-Phenylsulfonylvalerophenone. 0.0026M C16 (1.67) Col #1 145°C prod/std act 1.27 acet 0.71 acet(1.0M dioxane) 0.59 4-BB 1.16 4-BB(1.0M dioxane) 1.16 Table 57. Solvent Effects on Quantum Yield for S-Chloro- valerophenone. 0.00264M C16 (1.67) 0.10M pyr Col #1 145°C prod/std (acet) prod/std (4-BB) act benzene dioxane acetonitrile methanol 1.60 2.68 1.76 2.60 1.75 0.59 0.39 0.55 0.39 178 Table 58. Solvent Effects on Quantum Yield for 5-Chloro- 4'-methoxyva1erophenone. 0.025M C16 (1.67) 0.10M pyr Col #1 145°C acet/4-BB acetonitrile 18 methanol 22 Table 59. Solvent Effects on Quantum Yield for 5-Chloro- 4'-trifluoromethylvalerophenone. 0.00242M C16 (2.04) 0.10M pyr Col #1 145°C prod/std (acet) prod/std (4-BB) act 1.41 -- benzene 3.04 0.47 dioxane 0.99 0.18 acetonitrile 3.05 0.24 methanol 2.10 0.15 179 Table 60. Data for trans—4-Chloro-1,4-dimethy1-1—benzoy1- cyclohexane. 0.0119M C16, 0.0024MC19 (1.56) , 0.0028MC14 (2.3) Napth quencher, 0.05M pyr Col #1 145°C [Q] prod/std ¢°/¢ 0.0 1.62a(0.16)b --- 0.002 1.28 (0.11) 1.27a(1.41)b 0.004 1.20 (0.08) 1.35 (2.06) 0.008 0.95 (0.06) 1.70 (2.78) act 1.22 DMNC 3.97 Bzald 0.13 alcC 0.18 aRefers to DMBC. b Refers to benzaldehyde. cRefers to the assumed bicyclic alcohol. 180 Table 61. Data for trans-4—Bromo-1,4-dimethy1-1—benzoy1- cyclohexane. 0.0129M C16, 0.0134M C14 (2.3) Napth quengher Col #1 145 C [Q] prod/std ¢°/¢ o.ooa(o.00)b 0.265(0.52) --- 0.004(0.02) 0.145(0.60) l.83(1.17) 0.008(0.04) ’ 0.100(0.78) 2.65(1.53) -- (0.08) -- (0.92) -- (1.80) act 0.442 Bza1d(0.05M pyr) 0.025 Bza1d(0.10M pyr) 0.026 DMBC(0.05M pyr)c A[pyr] = 0.016 DMBC(0.10M pyr)° A[pyr] = 0.018 aRefers to benzaldehyde. bRefers to DMBC. C ¢DMBC determined by measuring the disappearance of pyridine . 181 Table 62. Data for cis-4-Bromo-1,4-dimethy1-1-benzoy1cyclo- hexane. 0.0143M C16, 0.015M C14 (2.3) Napth quencher Col #4 145°C [Q] prod/std ¢°/¢ 0.000 0.050 (0.040) --- 0.010 0.030 (0.014) 1.70 (2.85) 0.017. 0.023 (0.001) 2.17 (4.80) act 0.304 Bza1d(0.05M pyr) 0.092 DMBC (0.05M pyr) 0.110 Table 63. The 3-Thiobuty1propiophenone Sensitized Isomeri- zation of cis-1,3-pentadiene. Col #4 58°C [t—P].M 0.55Mc+t 1.0(act) 0.0522 -_- 1.0 0.0278 1.98 1.5 0.0221 2.42 2.0 0.0203 2.64 2.5 0.0167 3.22 182 Table 64. The 4-Thiobutylbutyrophenone Sensitized Isomeri- zation of cis-1,3-pentadiene. Col #4 58°C [c-pJ'l m‘l [t-P] M 0 55/¢ ' ' ' c+t 1.0(act) 0.0673 --— 1.0 0.0429 1.57 1.5 0.0392 1.71 2.0 0.0329 2.04 Table 65. The 3-Butylsu1finylpropiophenone Sensitized Isomerization of cis-1,3-pentadiene. Col #4 58°C [C-Pl'l 14'1 [t-P] M 0 55/¢ ' ' ' c+t 1.0(act) 0.0652 --- 1.8 . 0.0296 2.20 5.0 0.0163 3.98 Table 66. The 3-Buty1sulfonylpropiophenone Sensitized Isomerization of cis-1,3-pentadiene. Col #4 58°C [c-Pl-l M‘l [t-P] M 0 55/¢ ' ' ’ c+t 2.0(act) 0.0068 --- 125.0 0.0040 1.71 83.0 0.0016 1.45