:u. as! w fizfi; v I .0. .n . A. .1): Em: :1 . _. 3. 2% f... 5.1.3» .1: I: ... 1. “5X 51?; x0. .3 if: 1“ 4.1.2.. {‘9 A . ......~.......u..w... use". , J: 1.51M . ~31 T. V. C; 3 (19¢. This is to certify that the dissertation entitled INVESTIGATION OF STOICHIOMETRIC AND CATALYTIC B-C BOND FORMATION BY GROUP 9 TRANSITION METAL BORYL COMPLEXES presented by Jian-Yang Cho has been accepted towards fulfillment of the requirements for Ph . DL degree in Chemi stry - Maegan j" Major professor Date 1 (L’q .. 02- MS U is an Affirmative Action/ Equal Opportunity Institution 0- 12771 LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 c;/ClRC/DateDue.p65op.15 __...____ , - w INVESTIGATION OF STOICHIOMETRIC AND CATALYTIC B—C BOND FORMATION BY GROUP 9 TRANSITION METAL BORYL COMPLEXES By Jian-Yang Cho A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirement for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2002 ABSTRACT INVESTIGATION OF STOICHIOMETRIC AND CATALYTIC B-C BOND FORMATION BY GROUP 9 TRANSITION METAL BORYL COMPLEXES By J ian-Yang Cho C-H bond activation has attracted considerable attention since hydrocarbon feedstocks are ubiquitous. However, the catalytic functionalization of hydrocarbons still represents a long-standing challenge in homogeneous and heterogeneous catalysis. Since applications of arylboronate esters in cross-coupling chemistry are expansive, the transformation depicted below would have broad appeal. Ar-H + H-BXZ ——> Ar-BXZ + H-H In 1999, our group demonstrated the first thermal, catalytic aromatic borylation reaction catalyzed by Cp*Ir(PMe3)(H)(BPin). In order to understand this important transformation, detailed mechanistic studies were carried out. Through detailed investigations, a remarkably selective iridium catalyst system was discovered for aromatic borylation reactions. Borylations of various mono-substituted arenes provide essentially a statistical distribution of m- and p-C6H4(X)(BPin). This unique steric directing effect in aromatic borylation provides a complementary means for regioselective functionalization of aromatic compounds. For example, 1,3-disubstituted benzene rings are selectively borylated at the 5-position. With the exception of electron-deficient arenes, this is typically the least activated site towards aromatic substitution. Using this new methodology, a variety of arylboronic esters can be synthesized in a relatively simple and efficient way directly from corresponding arenes and pinacolborane or pinacol diboron catalyzed by the new iridium catalyst system. The new iridium catalysts tolerate the entire range of aryl halides, ether, and ester functionalities and selectively functionalize aromatic C-H bonds. This remarkable selectivity broadens the potential applications for aromatic borylation. Stoichiometric reactions of (PMe3)4Ir(BPin) and fac-(PMe3)3Ir(BPin)3 with arenes were examined. Both Irl and Ir"I boryl complexes can effect benzene borylation. However, their reactions with iodobenzene differ substantially under both stoichiometric and catalytic conditions. In order to further investigate the possibility that metal boryl complexes are intermediates in borylation reactions, several derivatives containing alkyl, aryl, or silyl ligands were synthesized and fully characterized. Preliminary mechanistic studies on the new iridium catalyst system were carried ”W catalytic out and presently a mechanism involving Ir!" and Irv intermediates in an Ir cycle is suggested. Correlations between the stoichiometric and catalytic reactions provided a deeper insight into the mechanism of aromatic borylation. T 0 my family and wife for their support and love iv ACKNOWLEDGMENTS First, I would like to thank my research advisor, Mitch Smith, for his guidance and assistance throughout my graduate career at Michigan State University. He taught me the qualities for pursuing the truth in science and gave me the opportunity to work on some exciting chemistry. Although the life for being a graduate student is tough, it equipped me with essential tools and knowledge to face future challenge. I will always appreciate that. I am thankful to Dr. Odom and Dr. Maleczka for their stimulating discussions during our joint group meetings and Dr. Jackson and Dr. Pinnavaia for serving on my guidance committee. My thanks also go out to Dr. Carl Iverson for his patient teaching and sharing of experiences during my first year here. I was very lucky to work with some talented postdoctoral associates, Dr. Man Kin Tse and Dr. Daniel Holmes. I learned a great deal from them. I will value the friendships I gained in my time at MSU. They are Chris Radano, Jim Ciszewski, Chen-Yu Yeh, Baixin Qian, J ie Fang, and Kuei-Fang Hsu. Their encouragement makes my graduate life more passable. Special thanks go to Jim and Kuei-Fang for their help in obtaining X-ray crystal structure. I also appreciate my best partners in Taiwan, Dr. Yih-Hsing Lo and Yi-Wei Chao, for their long-distance friendships. Finally, I would like to thank my best friend, my soulmate, and my wonderful wife, Mi-J in Chae, for her love, support, and sacrifice. Thank you for everything. I also would like to express my appreciation and love to my family in Taiwan: Dad, Mom, my elder brother, and my younger sister for their love, encouragement, and emotional support. I could not have gotten through this without them. My appreciation also goes out to my father and mother in-law in South Korea for their caring and support. Thank you so much. vi TABLE OF CONTENTS LIST OF TABLES .................................................................................... x LIST OF FIGURES ................................................................................. xii LIST OF SYMBOLS AND ABBREVIATIONS ............................................. xvii CHAPTER 1 INTRODUCTION C-H Bond Activation and F unctionalization of Hydrocarbons ........................ 1 Thermodynamics of Borane F unctionalization of OH Bonds ........................ 4 Stoichiometric and Catalytic Borylation Reactions ..................................... 5 Synthetic Routes to Arylboronic Esters ................................................... 9 CHAPTER 2 MECHANISTIC INVESTIGATION OF Cp*Ir(PMe3)(H)(BPin) CATALYZED BORYLATION REACTIONS Comparison between Cp*Ir(PMe3)(H)(BPin) and Cp*Rh(n4-C6Me6) Pre-catalyst Systems in Borylation of Various Substituted Arenes ................................. 12 Metathesis Reactions between Cp*M(PMe3)(Ph)(H) (M = Ir, Rh) and Pinacolborane in C6D6 ..................................................................... 18 Mechanistic Studies of The Original Iridium System ................................. 26 CHAPTER 3 CATALYTIC BORYLATION REACTIONS OF AROMATIC COMPOUNDS Screenings of Phosphine Ligands, Other Donor Ligands, and Metal Complexes for Catalytic Benzene Borylation ......................................................... 36 Borylation of Substituted Benzenes ...................................................... 42 vii Steric, Electronic, and Directing Effects in Aromatic Borylation ................... 48 Competition Reactions ..................................................................... 56 CHAPTER 4 SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF IRIDIUM BORYL COMPLEXES Synthesis and Characterization of an Irl Boryl Complex .............................. 58 Substitution and Oxidative Addition Reactions of IrI Boryl Complexes ............ 68 Synthesis and Characterization of an Ir"I Boryl Complex ............................ 74 Synthesis and Characterization of Novel Metal Boryl Complexes Containing Alkyl, Aryl, or Silyl Ligands ............................................................. 76 Oxidation Chemistry of IrI Complex with Boranes .................................... 93 CHAPTER 5 PRELIMINARY MECHANISTIC STUDIES OF THE IRIDIUM/PHOSPHINE CATALYST SYSTEM FOR AROMATIC BORYLATION Stoichiometric Reactions of Ir1 and Ir!" Boryl Complexes with Arenes ............ 96 Correlation between Phosphine Ligands and Catalytic Activity ................... 108 Stoichiometric and Catalytic Borylations of Iodobenzene .......................... l 10 Kinetic Isotope Effects in Aromatic Borylation ...................................... l 12 Mechanistic Discussions ................................................................. l 18 CHAPTER 6 EXPERIMENTAL General Considerations .................................................................. 126 Syntheses ................................................................................... 128 Screening Experiments .................................................................. 135 NMR Tube Reactions .................................................................... 137 viii Kinetic Isotope Effect Experiments .................................................... 143 Arylboronate Ester Syntheses ........................................................... 146 Competitive Borylation Experiments .................................................. 160 Kinetics Experiments ..................................................................... 162 Crystal Structure Determinations and Refinement ................................... 164 APPENDIX A Summary of Crystal Data and Structure Refinement ......................................... 167 APPENDIX B Derivation of Rate Expressions for Chapter 4 ................................................. 173 APPENDIX C Kinetic Details ..................................................................................... 176 BIBLIOGRAPHY ................................................................................. 1 79 ix LIST OF TABLES Table 1. Isolated yields (based on HBPin) and isomer distributions for catalytic borylation of aromatic hydrocarbons catalyzed by solutions of compounds 1 and 3 ...... 13 Table 2. Relative ratios of arylboronic esters for borylations of equimolar mixtures of substituted arenes catalyzed by compounds 2 and 3 ............................................ 17 Table 3. Selected bond lengths [A] and angles [°] for 14 ...................................... 34 Table 4. Summary of borylation of benzene with HBPin in the presence of 2 mol% pre- catalyst at 150 °C .................................................................................... 37 Table 5. Borylation reaction of benzene with HBPin in the presence of 2 mol% 13 and 2 mol% chelating phosphine ligand ................................................................. 38 Table 6. Borylation reaction of benzene with HBPin in the presence of 2 mol% 13 and nitrogen, oxygen, or sulfur containing ligands .................................................. 39 Table 7. Borylation of benzene with HBPin in the presence of various metal precursors (M) and ligands ...................................................................................... 41 Table 8. Ir-catalyzed aromatic borylations. Reactions are run in neat arene, [Ir] = 2 mol%, [P]:[Ir] = 2:1, and yields are reported for isolated materials ................................... 42 Table 9. Borylations of unsymmetrical 1,2- and 1,4-disubstituted arenes with HBPin in the presence of 2 mol% l3 and 2 mol% dmpe. Reactions run in neat arene, and yields are reported for isolated materials ..................................................................... 48 Table 10. Ligand repulsive energies (in kcal/mol) computed using the universal force field and A-values (in kcal/mol) for a variety of organic substituents ....................... 50 Table 11. Comparison of isomer distribution between the experimental values and the calculated values derived from pure steric effect (ER) .......................................... 52 Table 12. Borylations of mono-substituted arenes with HBPin in the presence of 2 mol% 13 and 2 mol% dmpe. Reactions run in neat arene. Isomer distribution is obtained from area ratios in GC-FID chromatograms ............................................................ 53 Table 13. Comparison of isomer distribution between the experimental values and the estimated values derived from selectivity in borylation of mono-substituted arenes ...... 55 Table 14. Relative ratios of arylboronic esters for borylations of equimolar mixtures of substituted arenes catalyzed by 2 mol% 13 and 2 mol% dmpe ................................ 56 Table 15. Selected bond lengths [A] and angles [°] for 17 .................................... 65 Table 16. Selected bond lengths [A] and angles [°] for 18 ..................................... 67 Table 17. Comparison of boryl resonances of complexes 15, 17, 22, 23 and 24 in 11B NMR spectra ......................................................................................... 73 Table 18. Selected bond lengths [A] and angles [°] for 25 ..................................... 76 Table 19. Selected bond lengths [A] and angles [°] for 28 ..................................... 88 Table 20. Selected bond lengths [A] and angles [°] for 29 .................................... 91 Table 21. Comparisons of XzB-Ir-PMe3mm, to Bpm bond distances of iridium boryl complexes ............................................................................................ 92 Table 22. Borylation reactions with HBPin in a molar ratio 1:1 mixture of C6H6/C6D6 or 1,3,5-C6D3H3 catalyzed by Ir' and Ir‘" sources at 150 °c, [Ir] = 2 mol%, [PMe3]:[Ir] 2 2:1 ................................................................................................... 116 Table 23. Stoichiometric borylation reactions of 18 and 25 with a molar ratio 1:1 mixture Of C6H6/C6D6 OI' 1,3,5-C6D3H3 at 150 0C ....................................................... 1 18 xi LIST OF FIGURES Figure 1. Various pathways discovered for the activation of OH bonds ..................... 4 Figure 2. Functionalization of hydrocarbons by transition metal boryl complexes under photochemical conditions ........................................................................... 6 Figure 3. Selective functionalization of alkanes by transition metal boryl complexes. . ....7 Figure 4. The first thermal, catalytic example of aromatic borylation reaction .............. 8 Figure 5. Reaction of BzPinz in pentane catalyzed by Cp*Re(CO)3 ........................... 9 Figure 6. Traditional and direct routes to arylboronic esters from aromatic hydrocarbons ......................................................................................... l 0 Figure 7. Resonance structures of ethyl benzoate and N,N-diethyl benzamide. . . . . . . . 1 6 Figure 8. Initial proposed catalytic cycle for catalytic functionalization of hydrocarbon C-H bonds ............................................................................................ 19 Figure 9. The reaction between Cp*Ir(PMe3)(Ph)(H) (5) and HBPin in C6D6 at 150 °C after around 37% conversion from Cp*Ir(PMe3)(Ph)(H) to Cp*Ir(PMe3)(H)(BPin): (a) 1H NMR spectrum of Cp* region; (b) lH NMR spectrum of aromatic region ................................................................................................. 20 Figure 10. The metathesis reaction between compound 5 and HBPin in CbDb at 150 °C ...................................................................................................... 21 Figure 11. Thermolysis of Cp*Ir(PMe3)(H)(BPin) (1) in C6D6 ............................... 22 Figure 12. 1H NIVIR and 3|P{1H} NMR spectra of the thermolysis of compound 1 in C6D6 ................................................................................................... 22 Figure 13. The reaction of compound 4 with HBPin in C6D6 at elevated temperature. . ..23 Figure 14. Plot of ln([4]t/[4]o) vs. time (s) for the reaction of compound 4 with [HBPin] = 0.551 M and [HBPin] = 1.103 M in C6D6 at 95 °C, respectively ............................. 25 Figure 15. Eyring plot for the reaction of compound 4 with HBPin in C6D6. ([4]o = 0.046 M; [HBPin]o = 0.551 M; T = 338.15 to 388.15 K, AH1= 25.6 kcal/mol and AS‘ = -53 en.) .................................................................................................... 26 xii Figure 16. Potential crossover products from pseudo double-labeling crossover experiment ............................................................................................ 27 Figure 17. Separation of compounds 2, 8, and 9 in a GC-MS chromatogram ............... 28 Figure 18. Pseudo double-labeling crossover experiment ..................................... 29 Figure 19-1. The chromatogram of the crude mixture from benzene borylation with pinacolborane in the presence of 10 mol% of compound 8 and 10 mol% of compound 9 ....................................................................................................... 30 Figure 19-2. The chromatogram of the crude mixture from benzene borylation with pinacolborane in the presence of 10 mol% of compound 8 and 10 mol% of compound 9 ....................................................................................................... 31 Figure 20. Borylation reactions of anisole with 20 mol% loading of compound 1 and compound 11, respectively ......................................................................... 32 Figure 21. ORTEP diagram of (MesH)Ir(BPin)3 (14). Thermal ellipsoids are shown at 25% probability ...................................................................................... 33 Figure 22. Benzene borylation with HBPin catalyzed by 2 mol% 14 and 4 mol% PMe; ................................................................................................... 34 Figure 23-1. GC chromatogram of borylation of 1,3—dichlorobenzene catalyzed by the Rh pre-catalyst 3 .......................................................................................... 46 Figure 23-2. GC chromatogram of borylation of 1,3-dichlorobenzene catalyzed by the Ir pre-catalyst 13 and dppe ............................................................................ 47 Figure 24. The calculated value of isomer distribution of the borylation of 1,4- C6H4(CI)(CF3) ....................................................................................... 51 Figure 25. The estimated value of isomer distribution of the borylation of 2- chloroanisole ......................................................................................... 54 Figure 26. Two potential catalytic cycles for aromatic borylation: (Left) involving Irv”I intermediates; (right) involving Ir"W intermediates ............................................ 59 Figure 27. Cyclometallation of tris(trimethylphosphine)neopentyliridium(I) complex. . .60 Figure 28. The reaction between mer-(PMe3)3Ir(BPin)(H)(Cl) (15) and KO’Bu .......... 60 Figure 29. Deborylhalogenation reaction between complex 16 and KO’Bu ................ 62 xiii Figure 30. ORTEP diagram of mer,cis-(PMe3)3Ir(BPin)2Cl (17). Thermal ellipsoids are shown at 25% probability .......................................................................... 64 Figure 31. Syntheses of mer,cis-(PMe3)3Ir(BPin)2Cl (17) and (PMe3)4Ir(BPin) (18). . ....66 Figure 32. ORTEP diagram of (PMe3)4Ir(BPin) (18). Thermal ellipsoids are shown at 25% probability ...................................................................................... 67 Figure 33. The reaction between compound 18 and dppe ...................................... 68 Figure 34. H2, R3SiH, and HX' oxidative additions to IrX(CO)(dppe) complexes. . . . . ....71 Figure 35. Catecholborane (HBCat) oxidative additions to IrX(CO)(dppe) complexes...7l Figure 36. Pinacolborane (HBPin) oxidative addition to compound 18 ...................... 72 Figure 37. Chlorocatecholborane (ClBCat) oxidative addition to compound 18 ........... 73 Figure 38. Synthesis of fac-(PMe3)3Ir(BPin)3 (25) ............................................. 74 Figure 39. ORTEP diagram of fac—(PMe3)3Ir(BPin)3 (25). Thermal ellipsoids are shown at 25% probability. All oxygen and carbon labels are omitted for clarity. Hydrogen atoms are also omitted for clarity ......................................................................... 75 Figure 40. Reported complexes containing a boryl ligand and a o-bound carbon ligand .................................................................................................. 77 Figure 41. The reaction between (PMe3)4Rh(Me) and BzCatz ................................ 78 Figure 42. The reaction between (PMe3)4Ir(Me) and HBPin in pentane ..................... 79 Figure 43. The resonance of the methyl group of fac-(PMe3)31r(Me)(H)(BPin) (26) in the 1H NMR spectrum ................................................................................... 81 Figure 44. The resonances of PMe3 groups of fac-(PMe3)3Ir(Me)(H)(BPin) (26) in the 1H NMR spectrum. The peaks denoted with an asterisk (*) are due to PMe3 and BPin resonances of compound 27 ........................................................................ 81 Figure 45. NOE experiments of compound 26 (Irradiation of the hydride resonance at -11.30 ppm) ........................................................................................... 82 Figure 46. NOE experiments of compound 26 (Irradiation of the Me resonance at 0.40 ppm) ................................................................................................... 83 xiv Figure 47. The resonances of PMe3 groups of fac—(PMe3)3Ir(Me)(H)(BPin) (26) in the 3|P{1H} NMR spectrum. The peaks denoted with an asterisk (*) are due to PMe3 resonances of compound 27 ........................................................................ 84 Figure 48. HETCOR experiment (1H, 31P) to correlate the resonances of PMe3 groups in the lH NMR spectra to those in the 3 lP{1H} NMR spectra .................................... 84 Figure 49. The reaction between (PMe3)4Ir(Me) with 9-BBN ................................ 85 Figure 50. The NMR reaction of (PMe3)3Ir(Ph) with HBPin in toluene-d3 at room temperature .......................................................................................... 87 Figure 51. ORTEP of mer—(PMe3)3Ir(BPin)(H)(Ph) (28). Thermal ellipsoids are shown at 25% probability ...................................................................................... 87 Figure 52. The reaction between (PMe3)4Ir(BPin) (l8) and HSiEt; .......................... 89 Figure 53. ORTEP of fac-(PMe3)3Ir(H)(BPin)(SiEt3) (29). Thermal ellipsoids are shown at 25% probability ................................................................................... 91 Figure 54. The reactions of Ir(PMe3)3(COE)(Cl) (30) with nitrogen-containing boranes including H[B(NH)2C6H4], H[B(NH)2C10H6] (HBDAN), and H[B(NMe)2C6H4] ......... 95 Figure 55. Thermolysis of 18 in C6136 at 150 °C ................................................ 97 Figure 56. Thermolysis of 18 in C6D6 at 150 °C: 1H NMR spectrum before thermolysis ........................................................................................... 98 Figure 57. Thermolysis of 18 in C6D6 at 150 °C: 1H NMR spectrum after thermolysis. Small peaks around 1.28-1.42 ppm and 1.57-1.60 ppm were not identified ................ 99 Figure 58. Plot of ln([18]t/[18]0) vs. time (min) for the thermolysis of 18 in C6D6 at 130 °C .................................................................................................... 100 Figure 59. Two potential pathways to account for the stoichiometric reaction between 18 and C6D6 ............................................................................................. 101 Figure 60. Plot of 1/k0bs vs. [PMe3] of the thermolysis of 18 in C6D6 in the presence of various concentrations of PMe3 .................................................................. 102 Figure 61. Our proposed mechanism for the thermolysis of 18 in C6D6. . . . . . . . ............103 Figure 62. The process of thermolysis of 25 in C6D6 at 150 °C ............................. 103 Figure 63. Concentration of each species relative to internal standard (CbMeo) vs. time (min) for the thermolysis of 25 in C6D6 at 150 °C measured by 1H NMR ................. 105 XV Figure 64. Plot of ln([25]t/[25]o) vs. time (min) for the thermolysis of 25 in CbDb at 150 °C ..................................................................................................... 107 Figure 65. The reaction of (PMe3)4Ir(H) (37) with HBPin and BzPinz ..................... 108 Figure 66. 1H, 11B, and 31P{1H} NMR spectra of the off-white precipitate from the reaction between 18 and C6H51 ................................................................... l l 1 Figure 67. Thermolysis of 25 in iodobenzene .................................................. l 12 Figure 68. The process of reversible formation of nz—arene complexes .................... l 13 Figure 69. Observed kH/kD in the activation of a 1:1 mixture of CbHo/CODO by the intermediate [Cp*Rh(PMe3)] ..................................................................... 1 14 Figure 70. A trisboryl complex [Ir((dtbpy)(COE)(BPin)3] isolated by Miyaura and co- workers .............................................................................................. 119 Figure 71. Possible mechanisms for Ir"I borylation reaction ................................ 120 Figure 72. Thermolysis of compound 28 in C6D6 at 50 °C ................................... 121 Figure 73. A putative mechanism for aromatic borylations catalyzed by iridium boryl complexes .......................................................................................... 123 Figure 74. Iridium tris(boryl) intermediate in borylation reactions ......................... 124 xvi BDAN BzPinz ClBCat COD COE dddd 6.11. E120 equiv. HBPin Hz LIST OF SYMBOLS AND ABBREVIATIONS Angstrom B(NH)2C10H6 bis(pinacolato)diboron, Me4C202B—B02C2Me4 chloro-catecholborane 1,5-cyclooctadiene cyclooctene pentamethylcyclopentadienyl, nS-C5(CH3)5 degrees Celcius doublet doublet of doublet of doublet of doublet deuterium entropy units diethyl ether equivalent facial gas chromatography hour pinacolborane, HBOZCZMe4 hertz infrared coupling constant xvii kcal kH/kD kobs mer min mL mmol mM mol MS NMR Ph PMe; PPh3 Pin rate constant temperature in Kelvin kilocalorie ratio of isotope effect on observed rate constant observed rate constant liter, generic ligand multiplet methyl, -CH3 meridional minutes milliliters millimole millimolar mole mass spectrometry nuclear magnetic resonance phenyl, -C6H5 trimethylphosphine triphenylphosphine pinacol, 1,2-02C2Me42' quartet singlet, seconds triplet xviii tetrahydrofuran delta, ppm for NMR spectroscopy change in enthalpy change in entropy ligand hapticity of number “n” microliters xix CHAPTER 1 INTRODUCTION C-H Bond Activation and Functionalization of Hydrocarbons The selective transformation of carbon—hydrogen bonds to other functional groups represents a long-standing challenge in homogeneous and heterogeneous catalysis, because C-H bonds are the most ubiquitous chemical linkages in Nature. Elucidating the requirements necessary to effect their cleavage or their transformation into other bonds is based on our fundamental understanding of their chemical reactivity. Over the past two decades, many examples of OH activation at transition metal centers were reported.1 It has been a topic of great interest to the organometallic chemists. Saturated hydrocarbons are major constituents of natural gas and petroleum, but there are very few practical processes for converting them directly to more valuable chemicals. The lack of reactivity of alkane C-H bonds can be attributed to their high bond energies (typically 90-104 kcal/mol) and very low acidity or basicity. Despite the fact that C-H bonds are more difficult than other types of linkages to cleave, such as CC] and C-Br, they are not completely inert. Alkanes have been known to undergo a number of solution and gas- phase reactions that involve free radicals as intermediates. They exhibit some preference for reaction of tertiary C-H bonds over primary or secondary. There are also some examples of alkane reactions with superacids2 or ozone3; however, they are usually very unselective. In recent years, considerable work with stoichiometric activation of C-H bonds suggests that homogeneous organometallic systems can overcome some of these selectivity problems.4 Many examples suggest that a regioselectivity pattern with (Le, 1° > 2° > 3°) can be obtained. Despite the success in this area, few systems are capable of subsequent substrate functionalization and regeneration of the metal species as required for catalytic turnover. Thus, developing a method not only to selectively activate but also functionalize C-H bonds of hydrocarbons has been a “Holy Grail” in synthetic chemistry.1C Selective catalytic hydrocarbon functionalization is not unknown. In fact, Nature performs ambient temperature alkane functionalization constantly, and sometimes with excellent selectivity, through the use of oxygenase enzymes. Those, which belong to the monooxygenase cytochrome P-4505 and methane monooxygenase6 (MMO) families, have received a considerable amount of attention. These enzymes catalyze the incorporation of molecular oxygen into alkane C-H bonds with the simultaneous loss of water and oxidation of NADPH or NADH. In the case of cytochrome P-450, the enzyme active site has been found to contain an iron porphyrin complex with a sulfur-bound cysteine, which mediates the cleavage of 02 to generate an iron-oxo complex. This oxo complex is considered to be the active oxidant of alkane C-H bonds. The efficiency of biological oxygenase systems has stimulated a significant body of research devoted to developing both structural and functional mimics designed to oxidize alkane.7 Pathways discovered for activation of C-H bonds include (i) oxidative addition of R—H to transition metal, (ii) 0 bond metathesis between M-R’ and R-H, 1,2-addition of R- H to M=X (X=O, NR, CR2), (iii) electrophilic activation in the reaction of M-X (X= halide, hydroxide, trifiate, etc.) and R-H to generate M-R and HX, and (iv) metalloradical activation. Oxidative addition reactions are typical pathways for late transition metal complexes. For example, Cp*Ir(PMe3)(H)2 (2, Cp* = nS-C5Mes) loses H2 under photo- irradiation to generate the reactive species “Cp*Ir(PMe3)”, which subsequently activate the OH bonds of hydrocarbon substrates to form a metal alkyl hydride complex.8 o-Bond metathesis occurs mostly in early transition metal complexes. These reactions usually result in the interchange of metal and hydrocarbon alkyl fragments.9 1,2-addition reactions involve the addition of an alkane to metal-nonmetal multiple bond.10 However, the scope of this type reaction and its potential for alkane functionalization remains unclear. Electrophilie activation reactions involve an electrophilic metal center, which attacks C-H bonds of alkanes to form functionalized alkanes directly. This type of reaction is usually carried out in a strongly polar medium such as water or anhydrous strong acid. One example in which the reaction between the Rh (II) porphyrin complexes and alkane C-H bonds with the involvement of free alkyl radical, generated through abstraction of a hydrogen atom from alkane by the Rh center is classified as metalloradical activation (Figure 1).H Systems that can selectively and catalytically functionalize the OH bonds of hydrocarbons are extremely rare. There are some examples of functionalization processes mediated by homogeneous transition metal complexes, including dehydrogenation of alkanes,l2 carbonylation of benzene,13 carbonylation of pentane,l4 and acceptorless dehydrogenation of cyclic alkanes by iridium complex with PCP type ligand.IS Mechanistic insight into the fundamental processes and solutions of potential problems towards to the functionalization of hydrocarbons have been studied extensively and well developed in many stoichiometric reactions. Currently the biggest challenge in this field is to develop better catalysts system to activate and functionalize the C -H bonds of hydrocarbons for practical applications. Oxidative addition LnM" + R-H ——> LnMX+2(R)(H) Sigma-bond metathesis LnMx-R + R'-H———> LnMLRH R-H 1,2-addition LnMX=Y + R-H —-» LnIyIX-T (Y = 0, NR, CR2) R H Electrophilie activation LnMx-Y + R-H LnMx-R + H-Y (Y=halide, hydroxide, triflate, etc.) Metalloradieal activation . ”I R-H (porphyrInIRh (R) {(porphyrinIRh"}2 -— 2 (porphyrin)Rh".—' . (porphyrin)Rh"'(H) Figure 1. Various pathways discovered for the activation of OH bonds. Thermodynamics of Borane Functionalization of C-H Bonds Since the importance of boryl complexes as proposed intermediates in the transition metal catalyzed functionalization of organic compounds, studies concerning the fundamental properties and reaction chemistry of transition metal boryl complexes have been initiated since early 19905. Transition metal-ligand covalent bond energies are important in understanding catalysis. However, there has been few data available for boranes and no thermochemical data for transition metal boryl complexes until 1994. In that year, Rablen and Hartwigl6 reported calculation of B-H and B-C bond dissociation energies (BDEs) for a series of boranes. From the established thermochemical and computational data of borane reagents, the reaction in equation 1 is essentially therrnoneutral. Moreover, from calculated BDE’s for B-H, C-H, and BC bonds, synthesis of aryl boronic esters directly from boranes and arenes should be thermodynamically feasible. CH4 + HBCat =————————- CHaBCat + H2 AH° =-2.1kcal/mol (1) Stoichiometric and Catalytic Borylation Reactions In 1995, Hartwig et al.17 reported the functionalization of hydrocarbons by the reaction of arenes and alkenes with (CO)5Mn(BCat), (CO)5Re(BCat), and CpFe(CO)2(BCat) under photochemical conditions (Figure 2). Although dehydrogenative borylation of benzene has not been observed previously before this report, dehydrogenative borylation of alkenes has been observed in catalytic chemistry. 1 8 (CO)5M—BCat T Q—BCat + HBCat + M2(CO)10 + H2 + Re3(CO)12H3 M = Mn 45% 10-20% M = Re 55% a Q ,.Fe-—-BCat ————> BC t + F +H 0C / hv, 1 h Q 3 DZ 2 87% Figure 2. F unctionalization of hydrocarbons by transition metal boryl complexes under photochemical conditions. Waltz and Hartwig” in 1997 reported selective functionalization of alkane at terminal position to produce alkylboronate esters, which are common reagents in organic synthesis. They found that photochemical reaction of Cp*Fe(CO)2(BCat’) (Cp* = C5Me5, Cat’ = 1,2,-O2C6H2-3,5-(CH3)2), Cp*Ru(CO)2(BCat’), and Cp*W(CO)3(BCat’) with a series of alkanes gives alkylboronate esters with functionalization of alkane exclusively at the terminal position (Figure 3). The boryl complexes are rare chemical reagents that react selectively at the terminal position of alkane to provide simple functionalized products. From their mechanistic analysis, they believe ligand dissociation is induced photochemically and that thermal reaction of the resulting intermediate occurs with alkanes. They pointed out that CpFe(CO)2(BCat) can functionalize arenes but not alkanes, presumably because of the presence of spz—hybridized C-H bond in the Cp ring, which is preferred to be fiinctionalized than alkanes. Therefore, they prepared a derivative of this complex with the spz-hybridized position blocked to solve the problem. Indeed, the elimination of all accessible sp2 positions on the metal boryl complex does account for the unusual reactivity observed. Reaction with pentane gave l-pentylboronate ester as the only functionalization product in 85% yield. Reaction with ethylcyclohexane gave (2-cyclohexyl)-1-ethylboronate ester as the only functionalization product in 74% yield. Interestingly, selectivity for the two terminal position of isopentane was good. Functionalization at the less hindered position occurred in 55% yield, whereas functionalization at the more hindered terminal position occurred in only 2% yield. 0C :4— /\/\ > MBCat' + HBCat' M= Fe 28% ~ 10% M = Ru 40% trace hv M: /\/\/BCat' 85% M OT O‘L Q9 Me BCat' och. MB, —— 74% W 0 Me > 55% 2% by O t O-BCat' 22% I \ BCat' + Figure 3. Selective functionalization of alkanes by transition metal boryl complexes. Hartwig et al.20 also examined the photochemical reaction of CpFe(CO)2(BCat) in a variety of mono-substituted arene solvents including C6H5(Me), C6H5(OMe), C6H5(Cl), C6H5(CF3), and C6H5(NMe2). They found the reaction of CpFe(CO)2(BCat) with substituted arenes resulted in formation of only meta- and para—substituted arylboronate esters for all substituted except anisole, which showed substantial amounts of ortho- substituted product. For N,N-dimethylaniline, it showed a preference for reaction at the para position. In 1999, Iverson and Smith21 reported the first thermal, catalytic aromatic borylation reaction to synthesize aryl boronic esters directly from arenes and pinacolborane by pre-catalyst Cp*Ir(PMe3)(H)(BPin) (Figure 4). They demonstrated the catalytic viability of equation 2 for the first time. :fi‘ Ir ''''' I o O I 17 mol/o Me3P/ \ BPin CsHe+H' >CHBPin+ H BP'” 150 °C,120h 6 5 2 53% Figure 4. The first thermal, catalytic example of aromatic borylation reaction. Ar-H * HBX2 --———- Ar-BX2 + H2 AH° ~- 3.1 kcal/mol (2) The findings was discovered in investigating stoichiometric B-C bond formation in reactions between Cp*Ir(PMe3)(Ph)(H) and HBPin in C6D6. They noticed the substantial quantities of arylboron products were produced from catalytic solvent activation. Aside from methane-to-methanol conversion,22 catalytic C-H functionalizations for unactivated hydrocarbons are extremely rare. Since applications of boronate esters in Miyaura-Suzuki cross-coupling chemistry are expansive,23 the demonstration of catalytic viability of equation 2 is significant. Later in the same year, Chen and Hartwig24 reported catalytic, regiospecific end- functionalization of alkanes by the rhenium complex, Cp*Re(CO)3, under photochemical conditions. They suggested that the terminal boronate esters are kinetic products, and the selective firnctionalization most likely results from a regiospecific reaction of the rhenium bis-boryl complex with the alkane primary C-H bond (Figure 5). 0. ,0 2.4-5? 0 *Re co :1: .343. + R-H ° p ( )3 : R-BPin + HBPin Q Q hv, CO, 25°C R = n-C5H11 Figure 5. Reaction of B2Pin2 in pentane catalyzed by Cp*Re(CO)3. Synthetic Routes to Arylboronic Esters The palladium-catalyzed cross-coupling reaction between organoboron compounds and organic halides or tn'flates provides a powerful and general methodology for the formation of C-C bonds. Arylboron reagents are typically synthesized in a multistep process: (1) halogenations of arenes to form aryl halides, (2) treatment of aryl halides with magnesium to generate their Grignard reagents, (3) reactions of Grignard reagents with trialkyl borate to give final corresponding arylboronate esters. Miyaura et al.25 have described a clever arylboronate ester synthesis where the generation of Grignard and lithium reagents is avoided by using palladium catalysts to effect the desired transformation from borane reagents and halogenated arenes. Since halogenated arenes required for these approaches must be synthesized from hydrocarbon feedstock, direct routes to the arylboron reagents from hydrocarbons are attractive (Figure 6). [Traditional Method] l X] M9 E“01:03 QH—{w wow» Ether - MgX(OR) [Miyaura et al.] Q 1 X1 "Pd", BzPinz or HBPin, Base Q H .___p X 1‘ BPin DMSO [ Direct Method I catalyst H + H-B(OR)2 H s B(OR)2 ’ 2 Figure 6. Traditional and direct routes to arylboronic esters from aromatic hydrocarbons. Several research groups26 including our group have achieved functionalization of hydrocarbons using a borane as an approach in stoichiometric and catalytic reactions under photochemical or thermal conditions. A milder reaction condition, higher turnover number, more efficient system is desirable. Furthermore, Knochel27 and co-workers recently reported the kinetic and thermodynamic aspects in OH bond activation by direct 10 borane-hydrocarbon dehydrogenation. On the basis of a better understanding of the role of transition metal boryl complexes in the organic transformation reactions, our ultimate goal is to develop a better catalytic system not only to activate but also functionalize hydrocarbon C-H bonds to well utilize ubiquitous hydrocarbon feedstock. 11 CHAPTER 2 MECHANISTIC INVESTIGATION OF Cp*Ir(PMe3)(H)(BPin) CATALYZED BORYLATION REACTIONS Comparison between Cp*lr(PMe3)(H)(BPin) and Cp*Rh(n4-C6Me6) Pre—catalyst Systems in Borylation of Various Substituted Arenes In 1999, our group reported the first thermal, catalytic example of aromatic borylation reactions to generate arylboronic esters directly from unactivated arenes and boranes by pre-catalyst Cp*Ir(PMe3)(H)(BPin) (1, Cp* = nS-CsMes).2' Subsequently, Hartwi g and co-workers reported alkane and arene borylations with the use of much more reactive Rh pre-catalyst, such as Cp*Rh(n4-C6Me6) (3).26a Given the broad utility of arylboronic esters in Pd-catalyzed cross-couplings with aryl and alkyl halides,23 we were curious as to the extent of regioselectivity and functional group compatibility for analogous transformations of substituted arenes.26b In an initial borylations of substituted arenes, 20 mol% solutions of 1 or Cp*Ir(PMe3)(H)2 (2) were dissolved with HBPin in neat arene solvents, and the reactions were run at 150 °C. Once borylation has commenced, 31P NMR spectroscopy indicated that compound 1 is the predominant Ir species in solution with small quantities of compound 2 present in each case. Analogous borylations were also performed using the more active pre-catalyst 3. After the borane was consumed, product ratios were determined by GC analysis of crude reaction mixtures. Isolated yields of isomer mixtures are reported in Table 1, and the product assignments were corroborated by comparisons to authentic samples. 12 Table 1. Isolated yields (based on HBPin) and isomer distributions for catalytic borylation of aromatic hydrocarbons catalyzed by solutions of compounds 1 and 3. Arene Product(s) % yield (parazmetawrtho), time'I % yield (parametazonho), timeb 13. :3 53, 120 he 92, 2.5 h“ :0 3 Q Q 91 (33.9:62.0:4.1),e 51 h 72 (32.5:62.7:4.9),/ 3.5 h I! 3 CD I‘/‘ / 0 <31 99 (33.3:66.7:0.0), 17 h 84 (33.3:66.7:0.0), 1.5 h PinB I“ / 0 Z 0 55 (19.5:79.0:1.6), 65 h 65 (25.4:66.9:7.6), l h PinB I“ / Z Z :3” 65 (42.9:54.9:2.1), 3.5 h E 3 W I“ / 0 I 3 Q N 52 (31.1:68.0:0.9), 142 h 67 (33.2:66.l:0.7), 2 h Pin «A? F3C F3 81,610h 86,3h Q BPin F3C F3 60,g 151 h 73,”4h Pin -- 41, 6 h N/_\ N/ BPin F F F F 81,’18h 41,/0.511 FQ~H FQBPM F F F F F -- 46,“ 0.5 h NEt2 : 1O | / / O C :DEI Pine}, \ OEt -— --(33.0:57.4:9.6),’"1h O _ O PinB/x NEtz -- 50(14.0:27.7:58.3), 0.511 " 20 mol% 1, generated in situ from compound 2 and HBPin at 150 °C. b 2 mol% 3 at 150 °C. ‘ 20 mol% l at 150 °C. d GC yield reported in reference 18c. ‘ < 1% of the isomer mixture is C6115CH2BPin. / 3% of the isomer mixture is C6H5CH2BPin. 8 3% of the isolated product is m-C6H4(Me)(CH2BPin). hp 12% of the product is m-C5H4(Me)(CH2BPin). ' 4% of the isolated product is isomers of C6F4H(BPin). ’ 16% of the isolated product is isomers of C6F4H(BPin). " Reaction was run in a 2:1 mixture of 1,3,5-C6H3F3zp-xylene- dlo. ’ C6HF3(BPin)2 (7%) and an isomer of C5H3F2(BPin) (6%). "' Products were not isolated. 13 In a simple assay of regioselectivity, toluene borylation primarily gave a statistical distribution of m- and p-C6H4Me(BPin). In order to address the reversibility of the borylation reactions, the catalytic borylation of C6D6 by HBPin in the presence of m- C6H4Me(BPin) was examined. Isomerization of m-C6H4Me(BPin) to p-C6H4Me(BPin) or generation of toluene would indicate that borylation is reversible. Under typical reaction conditions where C6D6 is converted to C6D5BPin, m-C6H4Me(BPin) does not isomerize and toluene is not eliminated. Thus, the borylation products are kinetically determined. It is noteworthy that arene C-H bonds are functionalized in the presence of weaker benzylic C-H bonds. A range of monosubstituted arenes was examined to determine whether reaction conditions would tolerate heteroatom substituents and to assess the generality for statistical meta/para substitution. The results in Table 1 indicate that meta/para ratios are predominantly statistical with the largest deviations occurring for N,N-dimethylaniline, which favors para substitution, and anisole, which favors meta substitution. For Rh- catalyzed reactions, the deviations were relatively small, but meta borylation of anisole is pronounced for Ir system. Since cumene gave a statistical distribution of meta and para borylation products, electronic effects are responsible for enhanced para selectivity for N,N—dimethylaniline. Benzylic activation of toluene increased for Rh (~3 % PhCH2BPin) versus Ir (~1 % PhCH2BPin). The aversion from borylation ortho to aromatic substituents suggested that selective borylation should be possible for 1,3-disubstituted arenes. We initially examined this possibility for 1,3-C6H4(CF3)2 and found exclusive borylation at the 5- position in reactions catalyzed by solutions of 1 or 3. For the borylation of m-xylene mediated by compound 3, benzylic activation increases significantly relative to that for 14 toluene. It appears that steric directing effects will extend to heterocycles as aromatic borylation of 2,6-lutidine occurs exclusively at the 4-position. Directing effects based on steric effects in aromatic substitution are uncommon, and electronic effects generally dictate the substitution pattern. For example, in the Friedel-Crafis alkylation of 3- chlorotoluene, no meta-substitution products are observed since Me group is an ortho- and para-directing activator and Cl is an ortho- and para-directing deactivator. Generally, selective meta functionalization is difficult and requires strong electron- withdrawing groups for electrophilic or —donating groups for nucleophilic aromatic substitution, respectively. The unique steric directing effect in aromatic borylation provides a complementary mean for regioselective functionalization of aromatic compounds. Fluorinated arenes were also tested for compatibility. II- or Rh-catalyzed borylation of C6HF5 gave C6F5BPin as the primary product. Similarly, compound 3 catalyzes borylation in a 2:1 mixture of 1,3,5-C6H3F3 and p-xylene-dlo to yield C6H2F3BPin as the major product. Attempts to prepare the di- and triborated compounds, CbHF3(BPin)2 and C6F3(BPin)3, from stoichiometric amounts of 1,3,5—C6H3F3 in p- xylene-dm yielded significant quantities of C6H3F2(BPin) (~60% of the borylated products). The selectivity for C-H activation is significant considering that Rh-catalyzed reactions of silane with CoHFs give C-F activation products exclusively.28 For arenes bearing ester and amide functionality, reduction of the carbonyl groups could potentially compete with aromatic borylation. Since HBPin reacts sluggishly in uncatalyzed hydroborations, selective aromatic borylations of aryl esters and amides seemed possible. Hence, catalytic borylations of ethyl benzoate and diethyl benzamide by 15 compound 3 were examined. In both instances, the reactions gave primarily aromatic borylation. For ethyl benzoate, meta/para borylation dominates with a modest increase in ortho borylation (p:m:o = 1.00:1.74:0.29), whereas diethyl benzamide gave 0- C6H4(C(O)NEt2)(BPin) as the major isomer (p:m:o = 1.00:1.98:4.17). The shift in substitution pattern is consistent with chelate-directed borylation at the ortho position. Since resonance structure B has a larger contribution for an amide relative to an ester (Figure 7), chelation of the amide oxygen to Rh or B in the catalytically active species is more favorable for the amide. The statistical metazpara ratio for the minor isomers suggests that chelate and sterically directed pathways compete. ph E ‘——> Ph/NE a E = NEt2, OEt A B Figure 7. Resonance structures of ethyl benzoate and N,N-diethyl benzamide. To probe the role of electronic effects, relative product ratios from catalytic borylations in equimolar mixtures of substituted arenes were determined (Table 2). Electron-deficient arenes are generally more reactive in both systems, and relative rate differences for Ir are slightly more pronounced than those for Rh. Ir-catalyzed borylation in neat N,N-dimethylaniline was extremely slow. Factors besides deactivation of the arene ring may be responsible for the reduced reaction rate; such as, possible coordination of the nitrogen lone-pair electrons of aniline to the Ir metal center, which 16 would block the active site around Ir. This may explain why cumene borylation in N,N— dimethylaniline/cumene mixtures was suppressed relative to borylation in neat cumene. Table 2. Relative ratios of arylboronic esters for borylations of equimolar mixtures of substituted arenes catalyzed by compounds 2 and 3. X—Q/YO XOBPin: Y_8Pin (Catalyzed by solutions of (Catalyzed by solutions of 2) 3) x = C133, Y = CH3 89.0:11.0 73.0:270 x = OCH3, Y = CH3 62.0:38.0 51049.0 x = N(CH3)2, Y = CH3 -- 40.8:592 x = N(CH3)2, Y = CH(CH3)2 310690 40.1 :59.9 A comparison of pre-catalysts l and 3 in borylations of various substituted arenes revealed that the Ir system was more selective toward arene C-H activation. Given the importance of selectivity in chemical synthesis, these findings spurred a detailed investigation of the original Ir system. 17 Metathesis Reactions between Cp*M(PMe3)(Ph)(H) (M = Ir, Rh) and Pinacolborane in C6D6 Fundamental understanding of hydrocarbon activation by “Cp*M(PMe3)” (C p* = C5Me5, M = Ir, Rh) has been studied extensively by Jones’ and Bergman’s research groups. In the iridium system, Bergmanla and co-workers showed that, upon irradiation, an excited state of Cp*Ir(PMeg)H2 is formed. This rapidly extrudes H2, and leaves behind the reactive, coordinately unsaturated intermediate “Cp*Ir(PMe3)”. The intermediate inserts into a C-H bond of R-H via a three-center transition state, and leads to the formation of Cp*Ir(PMe3)(R)(H) complexes. At the same period of time, Jones and co- workers studied the related rhodium complexeslb to determine the relative stabilities of the Cp*Rh(PMe3)(Alkyl)(H) and Cp*Rh(PMe3)(Aryl)(H). They found a slight kinetic preference for benzene over propane and overwhelming thermodynamic preference for benzene oxidative addition. For the Cp*Rh(PMe3)(Ph)(H) (4) complex, a reversible reductive elimination of benzene was found to occur at convenient rate upon heating to ~ 60 °C in C6D6 solvent, producing Cp*Rh(PMe3)(C6D5)(D). With a detailed picture of transition metal mediated C-H activation established by several research groups, we chose a Lewis acidic reagent, boranes, as an approach to convert activated alkyl and aryl groups to functionalized organic products. Our initial proposed catalytic cycle is shown below (Figure 8). 18 R—H + H—BXZ = R—BX2 + H2 AH°~-3.1kca|/mo| R—H HX, gar M / M= Ir, Rh M--.. M ...... / '1”sz Me P/ IR M°3P (III)\H 3 (|||)\H H—ex2 H2 _ ¢ , M ..... / "'H M 93': (|'|)\H H_BX2 R—BX2 Figure 8. Initial proposed catalytic cycle for catalytic functionalization of hydrocarbon C-H bonds. In the proposed catalytic cycle, the reactive unsaturated intermediate “Cp*M(PMe3)” can activate the C-H bond of hydrocarbons to form Cp*Ir(PMe3)(R)(H). Hopefully the metathesis reaction between HBX2 and Cp*Ir(PMe3)(R)(H) can release R- BX2 and generate Cp*Ir(PMe3)H2. Then the dihydride complex can react further with HBX2 to generate Cp*Ir(PMe3)(H)(BX2), followed by reductive elimination of HBX2 to 19 regenerate the reactive 16 electron intermediate “Cp*M(PMe3)” to complete the catalytic cycle. In order to test the viability of this proposed catalytic cycle, the metathesis reaction between Cp*M(PMe3)(Ph)(H) (M = Ir, Rh) and HBPin in C6D6 were examined. For the reaction between Cp*Ir(PMe3)(Ph)(H) (5) and HBPin in C6D6 at 150 °C, after around 37% conversion (based on 1H NMR Spectra) from Cp*Ir(PMe3)(Ph)(H) to Cp*lr(PMe3)(H)(BPin), the ratio between C6HsBPin and C6DsBPin was found to be around 1:23 (Figure 9). (a) .‘j Cp*Ir(PMe3)(Ph)(H) (5) Cp*Ir(PMe3)(H)(BPin) (1) I 2.20 " ' 2.10" ' “2.00“ "1.90” 1.80" A 1.70 (b) ‘ .I PthPin 5 PhBPin 5 '8.007 ”17.80 ' V 97.60 ' ' V7.40 ' '720 A A 7.00 Figure 9. The reaction between Cp*Ir(PMe3)(Ph)(H) (5) and HBPin in C6D6 at 150 °C after around 37% conversion from Cp*Ir(PMe3)(Ph)(H) to Cp*lr(PMe3)(H)(BPin): (a) 1H NMR spectrum of Cp* region; (b) 1H NMR spectrum of aromatic region. 20 Substantial quantities of arylboron products were produced from catalytic solvent activation. Hence, we conclude that catalytic borylation is much faster than metathesis reaction between compound 5 and HBPin in C6D6, even though the metathesis process is viable (Figure 10). ¥ CsHsBPln ! HBPin H2 g I Ira. - Ir ''''' 3.4 1. ll' """ I ' 1., + __Z__. "H / BPIn Me3P/ \HcfiHS HBP'” Me3P/ \ Meal” \H H (5) (1) Figure 10. The metathesis reaction between compound 5 and HBPin in C6D6 at 150 °C. Thermal borane elimination from Cp*Ir(PMe3)(H)(BPin) (1) was assessed by heating C6D6 solution of the pure compounds. After two weeks at 200 °C and one day at 280 °C, the formation of C6D5BPin was not detected (Figure 11). If HBPin is reductively eliminated from compound 1, the reactive intermediate “Cp*Ir(PMe3)” can activate C-D bond of C6D6 to form Cp*Ir(PMe3)(C6D5)(D), which can react further with HBPin through metathesis process to generate C6D5BPin. Since only small quantities of Cp*Ir(PMe3)(C6D5)(D) (6) were generated in the reaction afier prolonged thermolysis, the HBPin reductive elimination pathway is not kinetically competent to account for catalysis. Therefore, a pathway involving HBPin reductive elimination to generate an active catalyst can be eliminated. The final 1H NMR and 3 1P NMR spectra of the reaction are shown in Figure 12. The moderate quantities of BPin-O-BPin observed may result from the reaction between trace moisture and HBPin. 21 g12 2 weeks 1 day + excess CSDB > 280 °C A’ XCstBPIn lr"'u, ' / BP'" 200 °c M83P (1) H 2 /|r""IIBPin + "H20" 7‘ 2 /|r""uH + PinB—O—Bpln Me3P \ M63P \ H H (1) (2) Figure 11. Thermolysis of Cp*Ir(PMe3)(H)(BPin) (l) in C6D6. BPin-(l) BPin-O-BPin Cp*-(1) and Cp*-(2) PMe3-(1) \ lHNMR .' PMe3-(2) ' ‘ Cp*-(6) ~ I » 1 II 9: K2 A U. ‘1 ’L ‘1 22 . - 2.40 I I 2.20 T "2.00“ ‘ 1.80 ”001.6? "TKO ' “1.20 ‘ 1.00 I 3'P NMR f (1) I, l *1 (2) <6) : I '- WW‘WWWWWUWLLENVIW‘MWW: . , ; ,WWWWWWW -38 -39 -40 -41 -42 ~43 -44 -45 -46 -47 -48 -49 Figure 12. 1H NMR and 3|P{'H} NMR spectra of the thermolysis of compound 1 in C6D6. 22 The analogous reaction in the Rh system was also examined. Benzene reductive elimination from Cp*Rh(PMe3)(Ph)(H) (4) occurred before metathesis reaction took place. In the reaction of compound 4 with HBPin in C6D6 at elevated temperature, compound 4 reductively eliminated C6H6, followed by oxidative addition of HBPin to form Cp*Rh(PMe3)(H)(BPin) (7) (Figure 13). Simplified rate expressions can be derived by applying the steady-state approximation. Application of the steady-state approximation to Figure 13 gives the rate law and kobs expression shown in Equations 3 and 4. h'n,” = Rh ———-> Rh-u,” . Me3P/ \ CGH5 k4 Me3P/ HBPin MegP/ \ BP'“ H H + 4 [7] l l CeHs Figure 13. The reaction of compound 4 with HBPin in C6D6 at elevated temperature. - dl4] dt = k... [4] <3) MkflHBPM] k_1[CSH6] ""' k2[HBPIn] (4) kobs hkflHBPm] 1f k-1[CeH6] >> K2[HBPIn] kObS = k_1[CGH6] (5) 23 If k-1[C6H6] >> k2[H;BPin], kob, is expected to exhibit a first order dependence on [HBPin] as shown in equation 5. This is indeed the case as shown in Figure 14. The kinetic data are consistent with this proposed rate law. The kobs was measured from 65 to 115 °C, and activation parameters were determined from the temperature dependence of kobs. From the Eyring plot (Figure 15) over this temperature range, activation parameters were obtained: AH“t = 25.6 kcal/mol and AS: = -5.3 e.u. For a comparison, Jones and co- workers found that a reversible reductive elimination of benzene from compound 4 in C6D6 solvent at ~60 °C to form Cp*Rh(PMe3)(C6D5)(D) followed first-order kinetics over a 46-degree temperature range. From the Eyring plot of the first—order rate constants, they obtained the activation parameters for arene loss: AHI = 30.5 (8) kcal/mol and AS: = 14.9 (2.5) e.u. They stated that the positive value for the entropy of activation is consistent with the formation of an intact, dissociating benzene molecule in the transition state.29 In the reaction of compound 4 with HBPin in C6D6, the small negative value for the entropy of activation suggests that the transition state of the reaction is more ordered than the ground state. From the activation parameters established by Jones and co-workers, the kobs for benzene elimination at 75 °C is calculated to be 6.98 x 104 s'1 and the kobs for the reaction of compound 4 with IIBPin in C6D6 at 75 °C is 5 x 10'5 5“. Since the overall rate kobs for the reaction of compound 4 with HBPin at 75 °C (5 x 10'5 s") is smaller than kobs for benzene elimination at 75 °C (6.98 x 10“1 s"), the rate determining step must involve a process other than benzene elimination. This piece of evidence suggests that H-B activation is the rate-determining step in this reaction. 24 y = -0.0003X + 0.0015 R2 = 0.9994 -15 _ 1.103 M |n(l4]1/l4]o) y = —0.0006x - 0.0276 R2 = 0.9942 '5 T I I I I I I 0 1000 2000 3000 4000 5000 6000 7000 8000 time (s) Figure 14. Plot of ln([4]t/[4]o) vs. time (s) for the reaction of compound 4 with [HBPin] = 0.551 M and [HBPin] = 1.103 M in C6D6 at 95 °C, respectively. 25 L. o L. N ln(kobs/T) i; y = -12881x+ 21.102 -16 R2=0.9963 -18 0.0025 0.0026 0.0027 0.0028 0.0029 1rr Figure 15. Eyring plot for the reaction of compound 4 with HBPin in C6D6. ([4]o = 0.046 M; [HBPin]O = 0.551 M; T = 338.15 to 388.15 K, AH: = 25.6 kcal/mol and A81: -5.3 e.u.). Mechanistic Studies of The Original Iridium System Compound 1 was stable in benzene solution after prolonged thermolysis, which excludes the pathway involving the elimination of HBPin from Cp*Ir(PMe3)(H)(BPin) (1). Regarding the possibility of PMe3 dissociation pathway to generate Cp*Ir(H)(BPin), an analog of Hartwig’s proposed intermediate in the Rh system,268 we designed a pseudo double—labeling crossover experiment to probe the PMe3 dissociation pathway. Two labeled complexes Cp*Ir(P(CD3)3)(H)2 (8) and (C5Me4Et)Ir(PMe3)(H)2 (9) were prepared. If phosphine does dissociate from Cp*Ir(PMe3)(H)(BPin), we expect to see the 26 two crossover products, Cp*Ir(PMe3)(H)2 (2) and (C5Me4Et)Ir(P(CD3)3)(H)2 (10) (Figure 16). Surprisingly, those iridium complexes can be easily quantified by GC-MS (Figure 17). ;CH3 gCH3 : Il : lI ' lI : ll r ””””” + r"-a = I"... + r"-., / ”H / "'H / "’H / "H (0301313 \H MeaP \H (0301313 H MeaP \ H (8) (9) (8) (9) crossover products >_:1? i»»» -p ...l...tlb.t -rstirlr_ o .. «3 ._ .. ~ 2 _. .. .2 4 888 . . C m . . a .. n . _ . . . 889. I» awe.)— ~ I_;_\ . _ _ 2 . 88% Im w. W . . m _ I_.)..__\n. @5— ~ m 1 OOOOOW _ . _ fem .__ . 8802 A 2 ooooom_ I a 1.3.1.2089 . 825: _ 1&1 60:82:94.4 .Efiwoamfiofio $2.00 a E a use .w .N mcgomEoO mo :osfimmem .: 25w:— 28 Benzene borylation with pinacolborane in the presence of 10 mol% of compound 8 and 10 mol% of compound 9 was carried out (Figure 18). The reaction proceeded smoothly to generate C6H5BPin and H2 as products. From the chromatogram of the crude mixture (Figure 19) there was no crossover products observed. Crossover during catalytic borylation was minimal. Therefore, phosphine dissociation pathway is unlikely. 10 mol% 8 and 10 mol% 9 . HBPin + CBHB 150 °C ~* CeHsBPm + H2 Figure 18. Pseudo double-labeling crossover experiment. 29 2:5 65:. 00.3 00.w~ 00.: 00.02 00.3 00.3 pLELPEIPIIDPpPPbIIP.IPLrDPDDPrfflhzt’PDPII 00.2 00.2 00.: 00.3 00.0 .PEPIPPEIP'rht.’ IDDP-PIIIP O 1. A- 1 000000N n..- -_-A .. 000000v : 0000000 .. 0000000 .5 0000000 H 'H"T"Ifi. 823:2? .0 055088 00 £638 02 can 0 058088 00 o\6_oE 02 0o 8:880 20 E 0522:0935 ES, cosflbon ocowcon :80 2328 25.8 2: 00 280036030 2:. .70— 0.53.0 30 2:5 25 00.2 0900 00.00 002 00.2 093 00.3 00.2 00.m_ 002 .PbepP'vr-pnI—Prbbln,.b1int4.m.P pup-uhtbniPrpbbb} O 00000N \4 88$ I m m I..’..—_\n_ 02 _ m 000000 fo _ w . 88on I m m :91. 6 e 85%? .0 085088 00 $38 00 van 0 085088 .00 £108 0_ 00 88808 2: 8 082003880 2:3 scum—bop 2583 800 8388 25.8 05 0o ESwOEEOEo 2E. .Né— charm 31 However, added PMe3 strongly inhibited catalysis where HBPin was present. The finding raised possibility that small quantities of phosphine free Irv species could be active. Cp*IrH4-x(BPin)x species30 (where x = 1, 2) formed in the thermolysis of Cp*IrH4 (11) and HBPin and Cp*Ir(H)2(BPin)2 (12) are Ir analogs of other intermediates proposed by Hartwi g in the Rh system. Anisole borylation with 20 mol% loadings of compound 11 and 1 were compared (Figure 20). The isomer ratios for 11, o:m:p = 3:49:48; for 1, 03171:}? = 2:79:19. From this experiment, Cp*IrH4-x(BPin)x intermediates could be eliminated because the borylation regioselectivities for 11 and 1 differed substantially. OMe + HBP' = 2 '" 150°C o:m:p=2:79:19 20 mol% 11 PinB®OM H + ' 3 e * Q—OMe HBPII‘I 150 °C 2 o:m:p=3z49248 Figure 20. Borylation reactions of anisole with 20 mol% loading of compound 1 and compound 11, respectively. Exclusion of a simple phosphine dissociative pathway narrows the plausible catalysts to two choices: (1) Ir phosphine species arising from Cp* loss or (2) species where both Cp* and PMe3 have been lost. The latter possibility is intriguing in light of Marder’s synthesis of (116-arene)1r(BCat)3 complexes (where Cat = ortho-catecholate) from (Ind)Ir(COD) (13, where Ind = nS-C9H7, COD = 1,5-cyclooctadiene) and HBC at in 32 arene solvents.31 Using an analogous route, we prepared (MesH)Ir(BPin)3 (14, where MesH = 116-mesitylene) in 19% yield from compound 13 and HBPin in mesitylene solvent?“ Single crystals of 14 were grown from pentane at —30 °C and the structure was further confirmed by single—crystal X-ray crystallographic analysis. The molecular structure of 14 is shown in Figure 21 and the distance from Ir(1) to the center of the mesitylene ring is 1.880 A. Figure 21. ORTEP diagram of (MesH)Ir(BPin)3 (14). Thermal ellipsoids are shown at 25% probability. 33 Table 3. Selected bond lengths [Al and angles [°] for 14. Bond Distance [A] Bonds Angle [°] Ir(1)—B(1) 2.051(1) B(1)-Ir(1)-B(2) 80.6(4) Ir(1)-B(2) 2.021(1) B(1)-Ir(1)—B(3) 81.7(4) Ir(1)—B(3) 2.039(1) B(2)-Ir(1)-B(3) 83.8(4) Compound 14 reacts with benzene at 150 °C to produce Ir metal and three equivalents of C6H5BPin, but it does not catalyze C6H5BPin formation from benzene and HBPin. Thus, it appears that phosphines or related donor ligands are required for catalysis. Using the lability of the mesitylene ligand in 14, Ir phosphine species can be generated in situ from 14 and appropriate phosphines. Borylation of benzene with the use of 2 mol% 14 and 4 mol% PMe3 was found to be a viable pre-catalyst for aromatic borylation reactions (Figure 22). 2 mol% 14, 4 mol% PMe3 CeHe + HBPin = C5H5BPin + H2 150 ° , 15 h C 98% 60 yield Figure 22. Benzene borylation with HBPin catalyzed by 2 mol% l4 and 4 mol% PMe3. From the experiments discussed previously, we ruled out a H-B elimination pathway, PM63 dissociation pathway, and a pathway where the active species is generated from both Cp* and PMe3 loss. The remaining candidates for active species are iridium phosphine boryl complexes. Syntheses of some model complexes to examine their stoichiometric reactions with arenes and screening for different metal complexes 34 and ligands combination for catalysis would hopefully help to determine the identity of the active species and understand the mechanism of the aromatic borylation in this system. In this chapter, through detailed mechanistic studies of the original Ir pre-catalyst system, we suggest that active species are iridium phosphine boryl complexes. 35 CHAPTER 3 CATALYTIC BORYLATION REACTIONS OF AROMATIC COMPOUNDS Screenings of Phosphine Ligands, Other Donor Ligands, and Metal Complexes for Catalytic Benzene Borylation From mechanistic investigation of the initial Cp*Ir(PMe3)(H)(BPin) (1) pre— catalyst system, it was found that the use of 2 mol% (MesH)Ir(BPin)3 (14) and 4 mol% PMe3 was a viable pre-catalyst for aromatic borylation reactions. The low isolated yield of 14 hampered screening efforts and precluded practical applications despite dramatic improvement in catalytic activity. Hence, we sought alternative means for generating active catalysts. Because NMR indicated virtually quantitative generation of 14 from (Ind)Ir(COD) (13), in situ generation of active catalysts by phosphine addition to 13 was examined.32 This approach was successful and we were able to conduct systematic studies of the effects of different phosphine ligands, various donor ligands, and metal complexes on catalytic activity. First, as shown in Table 4 different ratios between PMe3 and “Ir” were examined for catalytic activity. The results showed that borylation rates were appreciable when [P]:[Ir] < 3:1 but decreased dramatically when [P]:[Ir] ratio equaled or exceeded 3:1 (Entries 2-5). Several other mono-dentate phosphine ligands including PEt3, P’Pr3, P'Bu3, PCy3, and PPh3 were also tested as ligands for the catalytic borylation of benzene giving moderate GC yields (Entries 6-10). 36 2 mol% 13 or 14, x mol% PR3 + HBPin > Qapm + H2 150 °C Table 4. Summary of borylation of benzene with HBPin in the presence of 2 mol% pre- catalyst at 150 °C.33 Entry Pre-catalyst [P]:[Ir] Ratio Reaction Time (h) Y(‘l’./eol)d 1 P112; 21 15 98 2 P152; 1 :1 5 87 3 Flt/1:3 2:1 18 87 4 Flt/1:3 3:1 57 0.4 5 Flt/1:3 4:1 20 O 6 1311;} 2: l 13 79 7 P3113 2:1 93 80 8 P’IBA‘U3 2:1 21 71 9 PC; 21 46 79 10 Pi>i13 2:1 58 69 Reactions run in neat benzene. GC yields based on HBPin. In addition to mono—dentate phosphine ligands, chelating bidentate phosphine ligands were examined (Table 5). A dramatic increase in catalytic activity and turnover numbers were observed for the bidentate phosphines. 37 n 2 mol% 13, 2 mol% RZP PR2 9: <::>8-BP01 + H2 A Table 5. Borylation reaction of benzene with HBPin in the presence of 2 mol% 13 and 2 mol% chelating phosphine ligand.33 . Pre-catalyst Temperature Reaction Time Yield Entry L'gand Loading (°C) (h) (%) 1 thP PPh2 2 mol% 150 2 95 //\ 2 thp PPh2 2 mol% 150 16 87 /\ 3 "I192P PMez 2 mol% 150 3 79 4 MezP PMe2 2 mol% 150 2 84* 5 CY2P PCY2 2 mol% 150 3 86 6 Q 2 mol% 150 0.8 78 PhZP PPh2 7 Ph2P Pth 2 mol% 100 7 88 /\ 8 MezF’ PMez 2 mol% 100 77 18 9 M6213 PMez 2 m0]% 100 31 96 10 Csz PCy2 2 mol% 100 96 86 /_\ 11 MezP PMez 0.2 mol% 150 9 85* 12 MezP PMe2 0.02 mol% 150 61 9O Reactions run in neat benzene. GC yields based on HBPin.*Isolated yield. Chelating phosphines as ligands substantially increased catalytic activity and TONS as highlighted for 1,2-bis(dimethylphosphino)ethane (dmpe) (Entry 12), where the effective TON of 4500 represents an improvement of more than 1000-fold over pre- catalyst Cp*Ir(PMeg)(H)(BPin) (1). Furthermore, the borylation reactions can be run at 38 100 °C with a reasonable rate by using 1,2-bis(diphenylphosphino)ethane (dppe) as the ligand (Entry 7). Besides phosphorous containing ligands, nitrogen, oxygen, and sulfur containing ligands were also screened for catalytic activity as shown in Table 6. 2 mol% 13, 2 mol%X x Q + HBPin A = Q—BPin + H2 Table 6. Borylation reaction of benzene with HBPin (0.7 M) in the presence of 2 mol% 13 and nitrogen, oxygen, or sulfur containing 1i ands. . Ligand Temperature Reaction Yield Entry L'gand Loading (°C) Time (h) (%) N 1 l \ 4 mol% 100 31 69 / S 2 (\ /7 4 mol% 150 15 8 IT N 3 M 2 mol% 150 0.2 85 —N N— 4 /_‘ ’_\ 2mol% 150 <1 85 5 MezNUNMez 2 mol% 150 6 52 (We 6 2 mol% 150 14 1 1 0M9 7 /0\/\O/ 2 mol% 150 14 7 8 2 mol% 150 1 6 9 2 mol% 150 2.7 73 10 2 mol% 150 16.3 66 11 2 mol% 100 1.5 86 Pre-catalyst Temperature Reaction Yield Entry Pre'cata'y“ Loadigg (°C) Time (b) (0/0) 12 2 mol% 100 1 82 13 _ _ 2 mol% 100 1 84 N N 14 / \ / \ 2 mol% 50 16 85 Reactions run in neat benzene. GC yields based on HBPin. 2,2’-bipyridine (bpy), 1,10-phenanthroline, 4,4’-di-tert-butyl-2,2’-bipyn'dine (dtbpy)26f were found to be good ligands for borylation (Entries 3, 4, 11, 12, and 13). For bpy, the borylation occurred at relatively low temperature (50 °C) at an appreciable rate, and good yield (Entry 14, 85% yield). Attempts to reduce pre-catalyst loading to 0.02 mol% yielded significant quantities of decomposition and PinB-O-BPin. The use of thiophene, veratrole, DME, and 2,2’-bithiophene were inefficient as ligands for borylation of benzene (Entries 2, 6, 7, and 8). Ligands containing both ”hard” N and “soft” P were also examined (Entries 9 and 10). Catalysts containing 2— (diphenylphosphino)-2’-(N,N-dimethylamino)biphenyl and 2-(dicyclohexylphosphino)- 2’—(N,N-dimethylamino)biphenyl did not exhibit any advantages in catalysis over chelating phosphine ligands, dppe and dmpe, or chelating nitrogen ligands, bpy, 1,10- phenanthroline, and dtbpy. N,N,N,N-tetramethylethylenediamine (TMEDA) performed marginally as a ligand (Entry 5, 52% yield). Besides 13, different metal complex precursors were investigated. Since the approach where we generated active catalysts in situ from (Ind)Ir(COD) (13) and phosphine ligands was successful, we were interested if the active catalysts can be 40 generated in situ from the precursor [Ir(COD)Cl]2, which is the starting material for the preparation of 13, or even IrC13-xH20. In addition, some of the related Rh complexes were also screened for catalysis. The results are summarized in Table 7. 1 or 2 mol% "M", 2 mol% fl . + HBPin 4' BP'” " “2 A Table 7. Borylation of benzene with HBPin in the presence of various metal precursors (M) and ligands. Entry Pre-catalyst Pizzjtflgfi Tempeglture Reactig; Time Yzzlp o 1 “32.22% 3 3 o 2 33 33 o . o 4 ”€33.33” 33 3 o s ”magpie: 33 -— o 6 (“33.2% 33 3 o v “(2.33% 33 33 Reactions run in neat benzene. GC yields based on HBPin. [Ir(COD)C1]z is an air-stable, commercially available iridium(I) compound and has been demonstrated to be a good catalyst system in conjunction with dmpe,33 bpy or dtbpy (Entries 1, 2, and 3).34 In contrast, IrCl3°xH20 and [Rh(COD)Cl]2 were poor metal precursors for catalytic borylation (Entries 4, 5, 6 and 7), and no borylation occurred for the [Rh(COD)Cl]2/dppe pre-catalyst system after 15 hours at 100 °C (Entry 5). 41 Borylation of Substituted Benzenes . . . . . . . 23 S1nce applications of boronate esters 1n cross-couplmg chem1stry are extenswe, a convenient way to expand the library of boronate esters is desired. With this goal in mind, borylations of a variety of arenes were carried out. The results are summarized in Table 8. Table 8. Ir-catalyzed aromatic borylations. Reactions are run in neat arene, [1r] = 2 mol%, [P]:[Ir] = 2: 1, and yields are reported for isolated materials.33 . . Temp Time Yield Entry Substrate Product Arene.HBP1n Catalyst QCL (h) (°/o) 1 O‘F /_E F 10:1 13/dppe 100 17 84 BPin 2 00' /__§ Cl 10:1 13/dppe 100 17 83 BPin 3 Oar /_\\ Br 10:1 13/dppe 100 17 90 BPin 4 O, -- 10:1 l3/dppe 100 60 -- 5 O" Q‘l 10:1 l4/dppe 100 57 77 BPin B F B F 6 r O ' Q 4:1 l3/dppe 100 14 81 BPin F F 7 OF FQBPin 4:1 13/dmpe 150 1 63 F F F PinB F 8 CF F—QBPin 1:5 13/dmpe 150 62 76 F PinB F 42 Temp Time Yield Entry Substrate Product ArenezllBPm Catalyst (°C) (h) (o/o) cu BPin 1:1.5 l3/dppe 100 14 89 c 11 c: CIE Br Br 10 Q DBPM 1:1.5 l3/dppe 100 17 92 Br Br BPin Me Mei—QM? 12:1 l3/dmpe 150 112 68 Cl 12 Q CI~Q~CI 4:1 l3/dmpe 150 39 76 CD Me CI BPin 33 QM pmaQMe 12:1 13/dmpe 150 10 85 Me Me 14 Q0 PinB‘QCI 9:1 l3/dmpe 150 12 98 Cl Cl 15* Q0149 PinBQOMe 1:3 13/dmpe 150 95 62 OMe OMe Isolated yields based on HBPin. *Reaction run in cyclohexane. 13: (Ind)Ir(COD); 14: (MesH)Ir(BPin)3; dppe: 1,2-bis(diphenylphosphino)ethane; dmpe: 1 ,2-bis(dimethylphosphino)ethane. Dramatic differences in chemoselectivities between Ir and Rh catalysts were found for halogenated substrates, where the Ir catalysts preferentially activate C-H bonds. Thus, good yields of mono- or triborated products of 1,3,5-trifluorobenzene were obtained by adjusting the arenezHBPin ratio (Entries 7 and 8). In contrast, previous attempts to effect multiple borylations of 1,3,5-trifluorobenzene with the use of Rh catalysts 3 led to increased defluorination.26b Ir-catalyzed borylations of 1,3- 43 dichlorobenzene and 1,3-dibromobenzene generate meta-functionalized products in high yields (Entries 9 and 10), whereas dehalogenation is the dominant pathway in Rh- catalyzed reactions.35 A dramatic example of the difference between Rh- and Ir-catalyzed borylation is shown in Figure 23 where the GC chromatograms of 1,3-dichlorobenzene borylations catalyzed by a Rh pre-catalyst and catalyzed by a Ir pre-catalyst are compared. In the Rh-catalyzed borylation, dechloronation is a competitive side reaction pathway; on the contrary, in the Ir-catalyzed borylation a clean single product was obtained. The finding that aromatic C-Br bonds survive in the Ir-catalyzed reactions contrasts Pd-catalyzed reactions of boranes and aryl bromides, where the C-Br bonds are converted to C-B or C-H bonds.36 Because aryl iodides have the weakest carbon-halogen bonds, they are susceptible to reductive cleavage by transition metals. Hence it is not surprising that the Ir catalysts generated from 13 are ineffective for the aromatic borylation of iodobenzene (Entry 4). However, iodobenzene and HBPin reacted smoothly to yield a mixture of C6H4(I)(BPin) isomers when active catalysts are generated from the Ir"I source, (MesH)Ir(BPin)3 (14), and dppe (Entry 5). With the success in borylating iodobenzene, Ir catalysts have been shown to be compatible with the entire range of aryl halides.33 We also demonstrated that cyclohexane can function as an inert solvent (Entry 15), which is useful in borylations of more valuable substrates. Ir catalysts selectively borylate symmetrical 1,2-disubstituted arenes including 0- xylene, 1,2-dichlorobenzene, and veratrole at the 4-position to give a single borylation product (Entries 13, 14, and 15). Symmetrical 1,4-disubstituted arenes can also be selectively borylated at the 2-position (Entries 11 and 12). Borylations of 1,4- disubstituted arenes proceed slower than the corresponding 1,2- and 1,3-disubstituted arenes presumably due to steric bulkiness of the two substituents in 1,4-disubstituted arenes. Attempts to borylate 1,3,5-trichlorobenzene led to unidentified decomposition species and PinB-O-BPin. Borylation of 1,4-C6H4(Br)(F) occurred selectively at the postion ortho to F most likely due to the different steric bulkiness between Br and F (Entry 6). 4S 9.85 2:21 -1 gpl .. 4 . i 6.. = . — u..— 4 .c 4 7— if. «Emmy/l . caml© .. com _o . .. cow 2 cam/l GIG . coo / x _ r cow fl ooofl _o samioimsa a 2a 1 H. 82 _o . 85622? .M “men—88-05 ad 2.: t3 moi—Emu ocoNconeoEomwé; .«o coca—boa mo EEwSmEoEo 00 .—.MN unsur— 46 3:5 65: E 3 S A: w o v p u t p i i p > n r i i n i h .. com 1 cow . ooo . cow _0 .. 002 mcE 62823.44 dame new 2 $338-65 : 2t 23 woNbS—wu ocoNCono._oEo€-m._ Co coca—bop mo EEQOEoEo DO .N-m~ 25w:— 47 Steric, Electronic, and Directing Effects in Aromatic Borylation From previous studies as shown in Table 8, borylations of symmetrical 1,2- and 1,4-disubstituted arenes give a single borylation product. In order to study the roles of steric and electronic effects on the selectivity for the borylation reactions, borylations of unsymmetrical 1,2- and 1,4-disubstituted arenes were compared. The results are summarized in Table 9. Table 9. Borylations of unsymmetrical 1,2- and 1,4-disubstituted arenes with HBPin in the presence of 2 mol% l3 and 2 mol% dmpe. Reactions run in neat arene, and yields are reported for isolated materials. Entry Substrate Product Distribution (°/o)* T839 22;! F I F | l F3C‘O—Cl 3C C 30 C 5 78 PinB BPin 11.82882 2 MeOOF MeO F MeO F 23 62 PinB BPin 66:93.4 3 MeOOCI MeO CI MeO Cl 23 62 PinB BPin 68.0:32.0 4 00m Cl Me Cl Me 44 88 PinB BPin 43.5:56.5 48 Entry Substrate Product Distribution (°/o)* Time “em (11) (%) 5 QCI PinBQ—OMe PinB—Q01 12 73 OMe 9' OMe 48.52515 6 GM PinBQ—OMe PinBQMe 12 77 OMe Me 0M8 36.3:63.7 7 QM PinBQMe PinBQCI 16 89 Cl 0' Me 62.2:37.8 * The product distribution was determined from the area ratio of each isomer in the GC chromatogram of the crude reaction mixture. One of the products from borylation of an unsymmetrical 1,2- or 1.4- disubstituted arene was independently synthesized according to the literature.25 Calculated ligand repulsive energies, ER, have been demonstrated in White and co-workers report37 to provide reliable steric parameters for ligands in organometallic systems. ER is defined as the amount of pure steric repulsion between a ligand and the prototypical molecular fragment to which it is bonded [Cr(CO)5] or [CpRh(CO)]. Other parameters such as Taft-Dubois steric parameter, B’s, and A-values38 are used as standard measures of steric effects in organic chemistry. However, the experimentally based measures, B’s and A-values, are a product of both steric and electronic effects. Therefore, ER values are applied here in order to assess the pure steric effect on borylation reactions. Selected ER values and A-values are listed in Table 10. 49 Table 10. Ligand repulsive energies (in kcal/mol) computed using the universal force field and A-values (in kcal/mol) for a variety of organic substituents. Substituent ER (kcal/mol) A-value (kcal/mol) F 0.28 0.25 Cl 1 0.53 Me 18 1 .74 OMe 37 0.75 CF3 44 2.5 If only steric effects are considered, we can calculate and predict the isomer distribution for borylation of unsymmetrical 1,4-disubstituted arenes and compare that to experimental data. Since, for a 1,4-disubstituted arene, both substituents have two ortho sites, the calculated isomer distribution ratio can be determined by using the relative ER values of the two different substituents. For present purposes, it was assumed that ortho substitution would preferably occur adjacent to the substituent with the smaller steric factor (ER). For example, the ER value for CF3 group is 44 and the ER value for C1 is 1; therefore, the isomer distribution for 1,4-C6H4(Cl)(CF3) borylation is expected to be 1:44 (22:97.8) for 1,3,4-C6H3(Cl)(BPin)(CF3) to 1,2,4-C6H3(Cl)(BPin)(CF3) on the basis of ER values as illustrated in Figure 24. The result differs from the experimental value of 11.82882 (Entry 1). Apparently, other factors are also involved in determining the isomer distribution. In order to determine if a trend exists for various unsymmetrically 1,4- disubstituted arenes, their borylation chemistry was investigated. The comparison of isomer distribution between the experimental values and the calculated values derived 50 from ER values for other unsymmetrical 1,4-disubstituted arenes are summarized in Table 11. F3C-Q'Cl ER value: 44 (CF3) 1 (Cl) substitution ortho to CF3 substitution ortho to Cl F3C Cl F3C CI PinB BPin Expected ratio: 1 : 44 = 2.2% : 97.8% Figure 24. The calculated value of isomer distribution of the borylation of 1,4- C6H4(C1)(CF3)- 51 Table 11. Comparison of isomer distribution between the experimental values and the calculated values derived from pure steric effect (ER). . 0 Calculated Entry Substrate Isomer erture(Ai) (from ER) (0/0) I . 1 F3C2©LC PinB BPin 2.2.97.8 11.8:882 MeO F MeO F M o F ’ 2 e «0 PM BPin 0.8.99.2 66:93.4 MeO CI MeO Cl M 0 Cl ' 3 e O. PinB BPin 2.6.97.4 68.01320 (:1 Me CI Me Cl M ' 4 ® 8 PinB BPin 94'7'5'3 43.5:56.5 The data summarized above suggest that the electronic effects and other effects such as chelate—directed effects of substituents also contribute to the resulting isomer distributions. It was therefore seemed of interest to evaluate borylations of unsymmetrical 1,2-disubstituted arenes to assess the electronic effects of substituents for these aromatic borylation reactions. In order to obtain a qualitative understanding of electronic effects, the experimental data are compared to the calculated isomer distribution data, which are derived from the selectivities observed in borylation of mono-substituted arenes (Table 12). 52 Table 12. Borylations of mono-substituted arenes with HBPin in the presence of 2 mol% 13 and 2 mol% dmpe. Reactions run in neat arene. Isomer distribution is obtained from area ratios in GC-FID chromatograms. Isomer Distribution Selectivity (garazmetamrtho) (°/o) (parametazortho) (“/o) P' B QOMe m GOMe 19.1:76.0:5.0 3211638242 P. B / O—Me m ><:\>_Me 31.6:67.1:1.3 48.1:51.0:1.0 P'nB , Om ' QC. 23.1:76.9:0.0 37.5:62.5:0.0 Arene Products The isomer distribution of borylation of each substrate including anisole, toluene, and chlorobenzene is determined from the area ratio of each isomer in the GC chromatogram. For present purposes, it was assumed that the response factors were similar because the molecules are isomers. Therefore, we deemed it unnecessary to make calibration curves for these screening experiments. In each mono-substituted arene, borylation can occur at either of two meta positions, two ortho positions, and one para position. The reported selectivities are obtained by dividing the GC isomer distribution for each isomer by the potential number of sites of borylation (e.g. for meta position divide by 2). Then the selectivity of para substitution is normalized to 1.00, and meta and ortho selectivities are based on the normalized value. For example, for anisole borylation the selectivity for parazmetazortho is 1.00:1.99:O.13 (= 32.1:63.8:42) and for chlorobenzene borylation the selectivity for para:meta:ortho is 1.00:1.67:0.00 (= 37.5:62.5:0.0). The method for calculating the estimated value of isomer distribution for unsymmetrical 1,2-disubstituted arenes is illustrated for the borylation of 2-chloroanisole (Figure 25). For isomer (A), the borylation occurs at the position meta to Cl and para to 53 OMe group, the estimated selectivity for isomer (A) is the sum of the two normalized selectivities of the mono-substituted arenes (1.67 and 1.00). For isomer (B), the borylation occurs at the position meta to OMe group and para to C1, the estimated selectivity for isomer (B) is the sum of the two normalized selectivities of the mono- substituted arenes (1.99 and 1.00). The estimated selectivity for borylation of 2- chloroanisole is obtained as 2.67:2.99, which is equal to 47.2:52.8. PinB. : :Cl 3:0 OMe PinB OMe (A) (B) 1,67 (meta to Cl) 1.99 (meta to OMe) + 1.00 (para to OMe) + 1.00 (para to Cl) 2.67 2.99 (A):(B) = 2.67:2.99 = 47.2%:52.8%. Figure 25. The estimated value of isomer distribution of the borylation of 2- chloroanisole. 54 Table 13. Comparison of isomer distribution between the experimental values and the estimated values derived from selectivity in borylation of mono-substituted arenes. . Observed Estimated Entry Substrate Isomer Mixture (%) (°/o) 1 Q0 PinB OMe PinB CI 48.5:51.5 47.2:528 OMe Cl OMe 2 QMe PinBQOMe PinBQMe 36.3:63.7 40.8:592 OMe Me OMe 3 QMG PinB Me PinB‘Q'C' 622:37.8 56.4:436 CI Cl Me From the comparison in Table 13, the experimental isomer distribution data observed for those three 1,2-disubstituted arenes are similar to the estimated ones. The small deviations between the observed values and estimated values might result from the different dielectric constants of these substrates since the borylation reactions were carried out in neat arenes solvents (8 (chlorobenzene): 5.69; 8 (anisole): 4.30; 8 (toluene): 2.38; 8 (2-chlorotoluene): 4.72; 8 (Z-methylanisole): 3.50; 8 (2-chloroanisole): N/A).39 The results for borylations of 1,2-disubstituted arenes give some indication of the electronic effects of various substituents. It is clear from assessing and comparing the electronic effects for borylation of 1,2-disubstituted arenes and the steric effects from the ligand repulsive energies, ER, that the methoxy group has a meta directing effect contributing to the isomer distribution in borylation of Z-methylanisole and 2- chloroanisole. 55 Competition Reactions To probe the role of electronic effects in the new Ir catalyst system (2 mol% 13 and 2 mol% of dmpe), relative product ratios from catalytic borylations in equimolar mixtures of substituted arenes were determined (Table 14). Table 14. Relative ratios of arylboronic esters for borylations of equimolar mixtures of substituted arenes catalyzed by 2 mol% 13 and 2 mol% dmpe. Equimolar Mixtures of Borylation Product Distribution Entry Substituted Arenes F3C F3C 1 b / Q QBPM/ Dem F3C F30 35:96.5 cu F3C c1 F3C 3 Q/ Q 9333/ CW CI F3C Cl F3C 48.52515 3 @_ flow} 4;)» /3.c—§}e. PinB Pin 51.92481 4 cn—Q—cu/F3C—QCF3 CI‘QCI /=ac~§—>~CF3 Pin Pin 993207 5 3—<}3/ 4} egg. / Q. PinB PinB 98.7:1.3 56 For borylation of 1,3-disubstituted arenes, steric effect directs borylation in the meta position. Therefore, in the competition reaction between m-xylene and 1,3- C6H4(CF3)2, the arene selectivity, 3.5:96.5 (Entry 1), is solely governed by the relative electronic nature of the two arenes. As mentioned previously, it was found that electron- deficient arenes borylate faster than electron-rich arenes. In the competition reaction between 1,3-C6H4(C1)2 and 1,3-C6H4(CF3)2, the arene selectivity was found to be 48.5:51.5 (Entry 2). Switching substrates from 1,3-disubstituted arenes to 1,4- disubstituted arenes, dramatically different arene selectivities were observed. In the competition reaction between p-xylene and 1,4-C6H4(CF3)2, the selectivity changed from 35:96.5 for the 1,3-disubstituted variants to 519248.] (Entry 3). Similarly, in the competition reaction between 1,4-C6H4(C1)2 and 1,4-C6H4(CF3)2, the selectivity is changed from 48.5:51.5 for 1,3-disubstituted variants to 99.3:0.7 (Entry 4). These results undoubtedly demonstrate that steric effects play a crucial role in the borylation of 1,4- disubstituted arenes. The results from screening experiments show that any changes in phosphine ligands, other donor ligands, or metal complexes have dramatic effect on the catalytic activity. Steric effect governs the regioselectivity of aromatic borylation. For borylation of symmetric 1,2-, 1,4- and symmetric or unsymmetric 1,3-disubstituted arenes, a single borylation product is obtained. In addition, for borylation of unsymmetric 1,2- or 1,4- disubstituted arenes, product distribution is the result of steric, electronic, and some type of chelate-directed effect depending on the substituents such as a OMe group. 57 CHAPTER 4 SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF IRIDIUM BORYL COMPLEXES Synthesis and Characterization of an IrI Boryl Complex In order to gain mechanistic insight into the Ir—catalyzed borylation reaction and establish the most likely overall mechanistic pathway, it is important to examine the viability and determine the rates of each individual stage. Halpem’s4O stepwise analysis of the mechanism by which Wilkinson’s catalyst catalyzes olefin hydrogenation provides a significant “take home lesson”; namely, that the identification of a dominant or detectable species in a catalytic system may lead to incorrect interpretations of the reaction mechanism. Only when kinetic and thermodynamic measurements define the role of the complexes along the actual reaction path can the mechanism be defined. A multi-step reaction is very complicated, and the dominant mechanism changes when the nature of the pre—catalyst, the ligand, and/or the substrate is altered. From a mechanistic standpoint, there are two potential catalytic cycles involving oxidative addition and reductive elimination from Irv"I and/or Irm/v intermediates with Irl and Ir'" boryl intermediates being the most likely C-H activating species in the Irw” llIN Ir and/or cycles, respectively as shown in Figure 26. 58 lllll IIIN LI Ph -H L HBPin H2 [LAT—BPin] H2 ;|r_BPin Ph‘H lr' L I . BPin lrlll H H - BPin L,.| m m L,_| _Ph Lli'ggm L, IIBPin L/:'_H '7 Ir L’lr L—;lr'—BPin er er Lillr-Ph BPin BPin H #1 3pm lrlll ”I \\ BPin 4 L, ' . L,, I H HBP' L ( llr-BPin .r_ _ . HBPin L’ Ph-BPin '” H Ph BPIn Figure 26. Two potential catalytic cycles for aromatic borylation: (Left) involving Ir'i'” [v . . Ir"I intermediates. intermediates; (right) involving From competition reactions of mono-substituted arenes, it was found that electron-deficient arenes borylate faster than electron-rich arenes. The general observed trend suggests that the iridium metal center is electron rich, which implies that Irl intermediates may be the active species in the catalytic borylation reactions. In order to examine that possibility, the Ir1 boryl complex, (PMe3)4Ir(BPin), was first synthesized to evaluate its stoichiometric reaction with arenes. Flood41 and co-workers’ studies on the mechanism of cyclometallation, i.e. oxidative addition of a C-H bond of a ligand to form a chelate complex, of tris(trimethylphosphine)neopentyliridium(I) compound showed that the mechanism involves direct, concerted oxidative addition and reductive elimination of the C-H bond 59 interconverting the square-planar Irl and octahedral Ir'” centers without PMe; dissociation (Figure 27). pit/19371 of + Me3CCH2Li ‘PMeg figé Me3P7lr‘-PMeg o = Me3P;-lr Me3P;!r Me3P 93111309340 C Me3P -—>V 0 °C Me3P pMe3 Figure 27. Cyclometallation of tris(trimethylphosphine)neopentyliridium(I) complex. At first, the 16-electron square planar complex, [(PMe3)3Ir(BPin)], the proposed key intermediate in the borylation reactions, was the target complex to be synthesized. Initially, we attempted to use a dehydrohalogenation route to synthesize [(PMe3)3Ir(BPin)] from the reaction between mer-(PMe3)3Ir(BPin)(H)(Cl) (15)42 and KO’Bu (Figure 28). However, there was no reaction at room temperature, and at elevated temperature the reaction proceeded slowly to generate a complex mixture of Ir products. Several other reagents including NaN(SiMe3)2 and 'BuLi were also used and these attempts were unsuccessful. Presumably, [(PMe3)3Ir(BPin)], a coordinatively unsaturated 16-electron square planar complex, is unstable, thus rendering its isolation and characterization difficult. BPin PM |,PMe3 f 83 Me3P;Ir—H + KO‘Bu Me3P7lr-BPin + ’BuOH 3» KCI Me3P ('31 Me3P (15) Figure 28. The reaction between mer-(PMe;)3Ir(BPin)(H)(Cl) (15) and KO’Bu. 6O In 1997, Marder, Norman, and co-workers reported the synthesis of a low valent, electron-rich, late transition metal boryl complex, (PMe3).4Rh(BCat),43 from the reaction between (PMe3)4Rh(Me) and BzCatz with MeBCat as byproduct. With this information in hand, we targeted (PMe3)4Ir(BPin), which is a coordinatively saturated l8-electron complex and is presumably more stable than the original 16-electron target. In 1982, Thorn and Tulip44 reported the preparation of (PMe3)4Ir(H) from the reaction of [(PMe3)4IrH2]Cl, which can be prepared at ambient temperature by purging dihydrogen gas into a THF solution of (PMe3)4Ir(Cl), and KO’Bu through a dehydrohalogenation route. A similar reaction was carried out to first synthesize [(PMe3)4Ir(H)(BPin)]Cl (16) by addition of HBPin instead of H2 to the THF solution of (PMe3)4Ir(Cl). This attempt was successful and compound 16 was prepared in 71% yield. The structure of 16 was assigned according to 1H, 1'B, and 31P{1H} NMR spectroscopy. In the 1H NMR spectrum, a doublet of quartet is observed in the hydride region at —13.16 ppm (ZJHP = 119.0 Hz, 18.7 Hz), indicating a trans relationship to a PMe3 group and a cis to three PMe3 groups. A singlet at 1.22 ppm integrating to 12 protons is assigned as the resonance of BPin group. There are two sets of doublets and one triplet corresponding to three different PMe3 groups at 1.61 ppm ((1, 2Jim = 8.2 Hz, 9H), 1.63 ppm ((1, 2JHp = 7.3 Hz, 9H), and 1.74 ppm (t, 2JHP = 3.4 Hz, 18H, two mutually trans PMe3). In the 3'P{‘H} NMR spectrum, there are three different phosphorous resonances. A broad peak is observed at -67.2 ppm (the PMe3 trans to BPin group), a quartet is observed at —60.1 ppm (ZJpp = 20.8 Hz, the PMe3 trans to hydride), and a triplet is observed at -55.2 ppm (ZJpp = 22.9 Hz, two mutually trans PMe3 groups). In the 11B NMR, a broad peak is observed at 32.8 ppm. Unfortunately, its reaction with KO’Bu proceeded through an undesired 61 deborylhalogenation pathway instead of dehydrohalogenation pathway to give (PMe3)4Ir(H) and 'BuOBPin as products (Figure 29). Presumably the formation of ’BuOBPin is thermodynamically favored. Since deborylhalogenation is a dominant reaction pathway, a process involving the synthesis of a diboryl complex in the first step, and the reaction with KO’Bu through the deborylhalogenation pathway to give a IrI boryl complex would be a possible synthetic route. Dehydrohalogenation KOtBU X Phgemggj 18 OH op * U _ KCI M6313" Ir BPine — PMe CI PMQ Cl HBPin,THF Hallie—3] Me3P- lr- PMe3 25 °C 2d = P-lr- , a s Me3P' y MeegPl BPin (16) PM83 ‘50—‘31 Me P-li‘PMe3 + 'BuOBPin 'KCl 3 .‘JPMea’ Deborylhalogenation Figure 29. Deborylhalogenation reaction between complex 16 and KO’Bu. mer,cis-(PMe3)3Ir(BPin)2Cl (17) was prepared from the reaction between B2Pin2 and (PMe3)4Ir(Cl) in a THF solution at 70 °C for one day. The complex was purified by recrystallization from a pentane solution at -30 °C to give spectroscopically pure compound 17 in 74% yield (Figure 31). The 1'B NMR spectrum shows two boryl resonances at 28.0 and 36.5 ppm and the 31P{'H} NMR spectrum exhibits a broad singlet at —51.4 ppm due to trans coupling to a boron nucleus (”B, spin 3/2, 80.4% natural 62 abundance, 10B, spin 3, 19.6% natural abundance)45 and a doublet at —41.1 ppm (ZJpp = 26.9 Hz). The catecholate analogue, mer,cis-(PMe3)3Ir(BCat)2Cl, was previously prepared via a different route by Dai and co-workers.46 Single crystals of 17 and B2Pin2 co—crystallized from pentane at —30 °C and the structures were established by X-ray crystallographic analysis. The molecular structure of 17 is shown in Figure 30. Selected bond distances and bond angles are given in Table 15. The molecular structure of 17 consists of an octahedral geometry with phosphines ligands in a meridional arrangement. The two BPin groups are cis to each other with the B(1)— Ir(1)—B(2) angle of 77.97(19). The boron trans to chloride has an Ir-B bond distance of 2.057(5) A, which is the same (within statistical error) as that in the compound mer- (PMe3)3Ir(BPin)(H)(Cl) (15) (2.054(3) A). The other Ir-B bond distance is 2.114(5) A. Among those Ir-P bond distances in complex 17, the two trans phosphine ligands have Ir- P distances of 2.328(1) and 2.320(1) A, respectively, which are very similar to the two trans Ir-P distances in compound 15 (2.3277(8) and 2.3147(8) A). Furthermore, the PMe3 ligand trans to BPin in compound 17 has a substantially longer Ir-P distance of 2.3921(13) A. This suggests that the BPin group has a larger trans influence than a PMe3 ligand. 63 Figure 30. ORTEP diagram of mer,cis-(PMe3)3Ir(BPin)2Cl (17). Thermal ellipsoids are shown at 25% probability. 64 Table 15. Selected bond lengths [A] and angles [°] for 17. Bond Distance [A] Bonds Angle [°] Ir(1)-B(1) 2.114(5) B(1)-Ir(1)-B(2) 78.0(2) Ir( 1 )-B(2) 2.057(5) B(l )-Ir(1)-P(3) 87.9(1) Ir(1)-P(3) 2.328(1) B(1)-Ir(1)-P(4) 164.3(1) Ir(1)-P(4) 2.392(1) B(1)-Ir(1)-P(5) 86.3(1) Ir(1)-P(5) 2.320(1) B(1)-Ir(1)-C1 102.7(1) Ir(1)-Cl 2.560(1) B(2)-Ir(1)-P(3) 96.8(1) B(2)-Ir(1)-P(4) 86.5(2) B(2)-Ir(1)-P(5) 97.2(1) B(2)-Ir(1)-Cl 179.3(2) P(3)-Ir(1)-P(4) 96.0(4) P(3)-Ir(1)-P(5) 163.33(4) P(3)-Ir(1)-Cl 8305(4) P(4)-Ir(1)-P(5) 9384(5) P(4)-Ir(1)-Cl 9288(4) P(5)-Ir(1)-Cl 8303(4) The reaction between compound 17 and KO’Bu in the presence of 2 equivalent PMe3 was carried out at ambient temperature to give (PMe3)4Ir(BPin) (18) in good yield (The product always contained a small amount of (PMe3)4Ir(H) (ca. 3% by 1H NMR) due to its considerable moisture sensitivity) (Figure 31). Complex 18 is fluxional in solution as evidenced by the appearance of a singlet at —57.5 ppm at 298.15 K (25 °C) in the 3 lP{‘H} NMR spectrum where it displays a doublet (-562 ppm, 2Jpp = 32.8 Hz, 3P) and a broad quartets (-57.7 ppm, 1P) at 213 K (-60 °C). The low temperature-limiting spectrum indicates a trigonal bipyramidal geometry with the BPin group occupying an axial site. 65 BPin . THF 1 p e (PMe ) Ir(Cl) : B Pin e .. ‘_ - 34 2 2 70°C,12h nggAr M913 CI (17) 74% yield KO’Bu,PMe3,THF 53"”. ea t > Me P-lr. + B OBP'n 17 25 °C, 90 min,- KCI 3 I Was U ' PM83 (18) 92% yield Figure 31. Syntheses of mer,cis-(PMe3)3Ir(BPin)2Cl (17) and (PMe3)41r(BPin) (18). Single crystals of 18 were grown from pentane at —30 °C and the structure was further confirmed by single—crystal X-ray crystallographic analysis. The molecular structure of 18 is shown in Figure 32. Selected bond distances and bond angles are given in Table 16. From the X-ray structure, the Ir(1)-P(1) distance of 2.334(2) A, is significantly longer than Ir(1)-P bonds of the other equatorial PMe3 ligands (Ir(1)-P(2), 2.287(2) A, Ir(1)-P(3), 2.282(2) A, and Ir(1)-P(4), 2.270(2) A). The result can be rationalized by assuming a large trans influence of BPin group. 66 Figure 32. ORTEP diagram of (PMe3)4Ir(BPin) (18). Thermal ellipsoids are shown at 25% probability. Table 16. Selected bond lengths [A] and angles [°] for 18. Bond Distance [A] Bonds Angle [°] Ir(1)-B(1) 2.147(9) B(1)-Ir(1)-P(1) 178.0(2) Ir(1)-P(1) 2.334(2) B(1)-Ir(1)-P(2) 85.1(2) Ir(1)-P(2) 2.287(2) B(1)-Ir(1)-P(3) 85.0(2) Ir(1)-P(3) 2.282(2) B(1)-Ir(1)-P(4) 82.2(2) Ir(1)-P(4) 2.270(2) P(2)-Ir(1)-P(3) 119.8(1) P(2)-Ir(1)-P(4) 119.1(1) P(3)-Ir(1)-P(4) 1 18.0(1) P(1)-Ir(1)-P(2) 95.8(7) P(1)-Ir(1)—P(3) 96.1(7) P(l)-Ir(1)-P(4) 95.8(8) 67 Substitution and Oxidative Addition Reactions of Ir1 Boryl Complexes (PMe3)4Ir(BPin) (18) underwent a variety of substitution and oxidative addition reactions. It reacted with dppe in C6D6 at room temperature for 3 days to give Ir(PMe3)2(dppe)(BPin) (19) as the major iridium containing complex as shown in Figure 33. The reaction resulted in the replacement of 2 PMe3 ligands with the chelating phosphine ligand, 1,2-bis(diphenylphosphino)ethane (dppe). The structure of 19 was assigned according to 1H, 11B, and 3 1P{1H} NMR spectroscopy. In the 1H NMR spectrum, a singlet at 1.10 ppm is assigned as the resonance of the BPin group. There is a triplet corresponding to two PMe3 groups at 1.33 ppm (ZJHP = 3.3 Hz, 18H) and a multiplet in the region of 1.92-2.18 ppm corresponding to four protons of the two CH2 groups on the backbone of dppe ligand. In the aromatic region (6.98-7.12, 7.16-7.28, 7.72-7.89, and 7.91-7.98 ppm) there are 20 protons corresponding to the protons for the four phenyl groups of the dppe ligand. In the HB NMR, a broad peak is observed at 38.8 ppm. In the 31P{1H} NMR spectrum, there are three different phosphorous resonances. A broad peak is observed at 46.1 ppm for the Pth trans to BPin, a doublet of triplet is observed at 39.1 ppm (J = 141.6 Hz, 13.4 Hz) for the Pth cis to BPin, and a doublet of doublet is observed at —58.9 ppm (J = 141.6 Hz, 26.8 Hz) corresponding to two PMe; groups. BPin BPin M P i ‘PMe3 * dppe Ph P i ‘PMea 2 PM 33 — “PM > 2 — r' + e3 1 ea 25 °C,3da s I Was PMe3 y PPh2 (1s) (19) Figure 33. The reaction between compound 18 and dppe. 68 Compound 18 reacted with HBPin in C6D6 to give mer,trans-(PMe3)3Ir(BPin)2(H) (20) as the initial predominant species, and this kinetic product gradually isomerized to fac-(PMe3)3Ir(BPin)2(H) (21) afier heating at 70 °C for 11 hours. The structure of 20 was assigned according to 1H, HB, and 3'P{1H} NMR spectroscopy. In the 1H NMR spectrum, a doublet of triplet is observed in the hydride region at —12.36 ppm (ZJHP = 117.0 Hz, 21.7 Hz), indicating a trans relationship to a PMe3 group and a cis orientation to two PMe3 groups. A singlet at 1.22 ppm is assigned as the resonance of the two mutually trans BPin groups. A doublet is observed at 1.49 ppm (J = 8.0 Hz, 9H) corresponding to the PMe3 trans to hydride and a triplet is observed at 1.74 (J = 3.4 Hz, 18H) corresponding to the two PMe3 groups, which are trans to each other. In the HB NMR, a broad peak is observed at 38.9 ppm. In the 31P{'H} NMR spectrum, there are two different phosphorous resonances. A triplet is observed at —59.6 ppm (J = 22.0 Hz) for the PMe3 trans to H and a doublet is observed at —50.8 ppm (J = 22.0 Hz) for the two mutually trans PMe3 groups. The structure of 21 was also assigned according to 1H, ‘1B, and 3'P{'H} NMR spectroscopy. In the 1H NMR spectrrun, a doublet of triplet is observed in the hydride region at —11.66 ppm (ZJHp = 118.1 Hz, 18.1 Hz), indicating a trans relationship to a PM63 group and a cis to two PMe3 groups. A singlet at 1.29 ppm is assigned as the resonance of the two chemically equivalent BPin groups. A virtual triplet is observed at 1.41 ppm corresponding to the two PMe3 groups trans to BPin and a doublet is observed at 1.58 ppm (J = 8.0 Hz) assigned as the PMe3 trans to hydride. In the 1]B NMR, a broad peak is observed at 38.6 ppm. The 31P{'H} NMR spectrum of 21 displays a triplet at —56.6 ppm (J = 22.0 Hz) for the PMe3 trans to H and a broad resonance at —61.8 ppm for the two PMe3 groups trans to BPin. 69 Eisenberg and co-workers47 reported that the oxidative addition of catecholborane (HBCat) to the Ir(I) cis-phosphine complexes IrX(CO)(dppe) (X = Br, I; dppe = 1,2- bis(diphenylphosphino)ethane) proceeds stereoselectively under kinetic control. In their original studies on H2, R3SiH, and HX' oxidative additions to IrX(CO)(dppe) complexes,48 they reasoned that both H2 and R3SiH add to the IrX(CO)(dppe) complexes as nucleophiles. The addition occurs over the OC-Ir-P axis because the 7r* orbital of CO is able to stabilize the developing transition state by reduction of electron density of the Ir dz2 orbital, thus minimizing the repulsive 4e' interaction between the filled dz2 and ob orbitals of H2 and R3SiH. On the contrary, hydrogen halides (HX') approach IrX(CO)(dppe) in aprotic media as electrophiles. Therefore, interactions that retain or enhance electron density at the metal center will favor addition along that pathway. Thus, HX' addition is preferred by bending the X-Ir-P axis because this pathway leads to an antibonding interaction between an occupied pz orbital of X and the dz2 orbital of Ir, thus enhancing the ability of Ir to donate electrons to the incoming electrophile (Figure 34).49 Since oxidative addition of catecholborane to IrX(CO)(dppe) complexes (X = Br, I) resembles HX' additions, the result implies that catecholborane approaches the metal center as an electrophile in accord with the view that the vacant B pz orbital can overlap with the filled dz2 of Ir (Figure 35). 7O (nucleophile) H-Y H Q Q . . Oé_'r—PPg—* + HY X—lripPfl 12* ”antibonding Y - H, R381 CO d22\v dzz p2 Ir-CO Ir-X (stabilized) (destabilized) 8‘ 8+ (electrophile) X'-H + HX' H32 ——-> .371. P3 F3: 00’ oc )'( 2 X' = Br, | dz [’2 Figure 34. H2, R3SiH, and HX' oxidative additions to IrX(CO)(dppe) complexes. (electrophile) (E E) 6+ 8‘ 5kg); 0 B, ,o .P H P 33:»? .3231: _. _. .,1_.2 oc :31”? oc ,2 x, P 00 Figure 35. Catecholborane (HBCat) oxidative addition to IrX(CO)(dppe) complexes. Pinacolborane (HBPin) oxidative addition to compound 18 gives complex 20 as the kinetic product with the two BPin groups trans to each other. The result can be rationalized by assuming that pinacolborane (HBPin) approaches the Ir metal center as an electrophile in accord with the results for HBCat. Furthermore, the backbonding interaction between the filled Ir metal dz2 orbital and the empty B pz orbital, as HBPin 71 adds in the P-Ir-B plane, provides additional stabilization to the five-coordinate species, as shown in Figure 36. (kinetic product) (thermodynamic product) BPin . Bp PMe3 1 PMe3 3 HBPm _]_{]e e3 11312.11 Me3P- 'r'PMe3 W Me;1,P-;lr-70 °C MegP: lr- BPin (18) (20)n (21) - PMR . HBP/ Me3P- lr- PBPein M6313! d22 . M93 dzz op 33 Me3P’ 3. empty 9: Me3P empty Pz M93P( M83P; 08% g 958 8%H antibondin empty Pz g/I C5) (electrophile) Figure 36. Pinacolborane (HBPin) oxidative addition to compound 18. The reaction of 18 with Chlorocatecholborane (ClBCat) proceeded at room temperature to give mer-(PMe3)3Ir(BPin)(BCat)(Cl) (22) as the major product with the BPin group trans to the chloride ligand as shown in Figure 37. The geometry of 22 was assigned according to 1H, HB, and 31NH} NMR spectroscopy. In the 1H NMR spectrum, a singlet at 1.18 ppm is assigned as the resonance of the BPin group. There is a doublet at 1.29 ppm (J = 7.3 Hz) corresponding to the PMe3 trans to BCat and a triplet at 72 1.47 ppm corresponding to two mutually trans PMe3 groups. An AA'BB' pattern is observed at 6.84 and 7.19 ppm in the aromatic region, which corresponds to the four protons of the catecholate. The HB NMR spectrum displays two well-separated resonances at 28.1 and 41.6 ppm, respectively. By comparing the resonances of boryl groups in compound 15, 17, mer—(PMe3)3Ir(BCat)(H)Cl (23),42 and mer,cis- (PMe3)3Ir(BCat)2Cl (24)46 as summarized in Table 17, we assigned the resonance at 28.1 ppm as the BPin group and the resonance at 41.6 ppm as the BCat group. The 3'P{IH} NMR spectrum of 22 displays a broad peak at -56.2 ppm assigned as the PMe3 trans to BCat and a doublet at —39.3 ppm (J = 29.3 Hz) corresponding to two mutually trans PMe3 groups. BPin BPin | .PMe3 * ClBCat, - PMe3 |‘PMe3 Me3P—jr'PMe3 25 °C 3 d > Me3P;lr-BCat , a s PMe3 y Me3p Cl (18) (22) Figure 37. Chlorocatecholborane (ClBCat) oxidative addition to compound 18. Table 17. Comparison of boryl resonances of complexes 15, 17, 22, 23, and 24 in 1]B NMR spectra. Complex 15 17 22 23 24 BPin trans to Cl (ppm) 28.5 28.0 28.1 -- -- BCat trans to Cl (ppm) -- -- 32.8 32.6 BPin trans to PMe3 (ppm) -- 36.5 -- -- -- BCat trans to PMe3 (ppm) -- -- 41.6 -- 41.7 73 Ill Synthesis and Characterization of an Ir Boryl Complex The Ir'” boryl complex, fac-(PMe3)3Ir(BPin)3 (25) can be easily prepared in essentially quantitative yield by addition of a slight excess of PMe3 to (MesH)Ir(BPin)3 (14) in benzene solution at ambient temperature (Figure 38). Q . BPin | PMea (5 equw.). CoHe l BPin ...,,,, . = Me P-lr'-BPin Ir BP'” 25 °C, 30 min Mea3p' I PinB BF," PMea 1 (14) (25) 99% isolated yield Figure 38. Synthesis of fac-(PMe3)3Ir(BPin)3 (25). Crystals suitable for X-ray analysis of 25 were grown from pentane at —30 °C. The molecular structure of 25 is shown in Figure 39. Selected bond distances and bond angles are given in Table 18. The complex adopts a facial geometry with all three PMe3 ligands trans to the BPin groups. The 11B NMR spectrum shows only one boryl resonance at 36 ppm and the 31P {'H} NMR spectrum shows one broad singlet at -64 ppm due to a trans coupling to the boron nucleus. A similar compound, fac-(PEt3)3Ir(BCat)3, was previously prepared by Marder and co-workers via the displacement of the mesitylene ligand of (MesH)Ir(BCat)3 with 3 equivalent of PEt3.31 fac-(PEt3)3Ir(BCat)3 has similar spectroscopic properties as compound 25. In the HB NMR spectrum, a broad peak was observed at 44.7 ppm corresponding to the three BCat groups and in 31P{'H} 74 NMR spectrum, a broad peak was observed at —32.1 ppm corresponding to the three PMe3 groups. Figure 39. ORTEP diagram of fac—(PMe3)3Ir(BPin)3 (25). Thermal ellipsoids are shown at 25% probability. All oxygen and carbon labels are omitted for clarity. Hydrogen atoms are also omitted for clarity. 75 Table 18. Selected bond lengths [A] and agles [°] for 25. Bond Distance [A] Bonds Angle [°] Ir(1)-B(1) 2.106(2) B(1)-Ir(1)-B(2) 82.3(7) Ir(1)-B(2) 2.148(2) B(1)-Ir(1)-B(3) 79.4(6) Ir(1)-B(3) 2.082(2) B(2)-Ir(l )-B(3) 81 .9(7) Ir(1)-P(1) 2.352(4) B(1)-Ir(1)-P(1) 169.2(5) Ir(1)-P(2) 2.353(4) B(1)-Ir(1)-P(2) 91 .5(5) Ir(1)-P(3) 2.347(4) B(1)-Ir(1)-P(3) 89.6(5) B(2)-Ir(1 )-P(1 ) 90.8(5) B(2)-Ir( l )-P(2) 90.9(5) B(2)—Ir(1)—P(3) 169.6(5) B(3)-Ir(l)-P( 1) 91 .4(5) B(3)-Ir(l)-P(2) 169.0(5) B(3)-Ir( l )-P(3) 90.1(5) P(1)-Ir(1)-P(2) 97.1(2) P(1)-Ir(1)-P(3) 962(2) P(2)-Ir(1)-P(3) 95 .9(2) Synthesis and Characterization of Novel Metal Boryl Complexes Containing Alkyl, Aryl, or Silyl Ligands In transition metal catalyzed hydroboration of unsaturated hydrocarbons and borylation reactions of alkanes and arenes, an intermediate containing a boryl ligand and a o-bound carbon ligand has been proposed. Several theoretical studies also support the interrnediacy of such metal complex. However, only a few of this type of metal complexes have appeared in literature (Figure 40). One of them, reported by Crabtree50 and co-workers in 1993, was an unprecedented IrIV boryl complex, [Ir(PMe3)3(biphBF)Cl]+I'. Knorr and Merola42 reported the preparation and characterization of mer-(PMe3)3Ir(BCat)(trans-{CO2CH3}=CH {CO2CH3})(C1) in which the vinyl group arises from insertion of dimethyl acetylenedicarboxylate into the Ir-H 76 bond of mer-(PMe3)3Ir(BCat)(H)(Cl). Another case is the report by Roper, Wright, and co-workers of the synthesis of cis- and trans-[Os(BCat)(o-tolyl)(CO)2(PPh;1)2].5’I The complex with o-tolyl and BCat groups cis to each other was unstable at room temperature. o-tolleCat was slowly eliminated from the complex at room temperature in i benzene solution to give the orthometallation product, [Os(C6H4PPh2)(H)(CO)2(PPh3)]. On the other hand, the complex with o-tolyl and BCat groups trans to each other was stable under reflux in benzene for 2 hours. Their observations indicate that the requirement for facile reductive elimination is that the aryl and BCat ligands be adjacent to one another. 0 PMe3 II ,PMe3 |r\\ PMe3 CI -03” l PPh3 oc,, ,,,,, I ..... .R o 00/ l S\B\’O PPh3 0D R = o-tolyl o, o T 0PMe3 H M63P7llr'fi¢ SiPh3 > H > CH3 based on X-ray structural data. A good trans-influence ligand could weaken the bond between the metal and the trans ligand. This is a thermodynamic effect. The study shows that silyl group has stronger trans influence than hydride. Moreover, in the computational study by Sakaki and co-workers,56 they noted that the trans influence of the boryl group is stronger than the very strong trans influence of silyl group. We deemed it of interest to establish experimentally the stronger trans influence of a boryl group than that of a silyl group. Thus, we decided to explore the reaction between (PMe3)4Ir(BPin) (18) and HSiEt3. The reaction proceeded slowly at room temperature over 2.5 days to give fac- (PMe3)3Ir(H)(BPin)(SiEt3) (29) as the only observed product (Figure 52). 88 BPin [I’Mea , 1 25 °C. 2.5 days .H Me3pl'1r-PMe3 + 14311513 0 H > Me3ls;1r-BPin + PMe3 M P e . (18) fee-(29) 86% yield Figure 52. The reaction between (PMe3)4Ir(BPin) (18) and HSiEt3. The product was then recrystallized from a concentrated pentane solution at —30 °C to give 83 mg of colorless crystals in 86% yield. The structure of 29 was established by 1H, llB, 3'P{1H} NMR spectroscopy. In the 1H NMR spectrum, a doublet of triplet is observed in the hydride region at ~12.30 ppm (hydride, 1H, 2JHp = 117.0 Hz, 17.0 Hz), indicating that it is trans to a PMe3 group and cis to two PMe; groups. A singlet at 1.29 ppm is assigned as the resonance of the BPin group. There are three sets of doublets corresponding to three different PMe3 groups at 1.25 ppm (2.11“) = 6.7 Hz, PMe3 trans to BPin), 1.37 ppm (ZJHp = 7.3 Hz, PMe; trans to SiEt3), and 1.46 ppm (ZJHp = 7.6 Hz, PMe; trans to Me). There is no symmetry about the Ir metal center, which renders the molecule chiral. Thus, the CH2 groups of SiEt3 are diastereotopic and should exhibit complex coupling patterns. Indeed, three diastereotopic protons of the CH2 groups of SiEt3 are observed from 0.84 to 0.95 ppm and there are twelve protons in the region of 1.35-1.45 ppm which include the other three diastereotopic protons of the CH2 groups of SiEt; and nine protons from the CH3 groups of SiEt3. In the 3]NH} NMR spectrum, there are three different phosphorous resonances. A broad peak is observed at -66.4 ppm corresponding to the PMe3 trans to BPin group, a doublet of doublet is observed at —64.2 ppm (2Jpp = 31.3 Hz, 19.8 Hz, PMe3 trans to SiEt3), and another doublet of doublet is observed at 89 —58.1 ppm (ZJpp = 19.8 Hz, 19.8 Hz, PMe3 trans to H). The assignment of the PMe3 groups in the 1H NMR spectrum is established by one-dimensional Nuclear Overhauser Effect (NOE) experiments and the assignment of the PMe3 groups in the 3‘P{'H} NMR spectrum is based on two-dimensional selective decoupling experiments ('H, 31P) to correlate the resonances of PMe3 groups in the 1H NMR spectrum to those in the 3 lP{1H} NMR spectrum. Complex 29 adopts a facial geometry with BPin, SiEt3, and H groups all trans to different PMe3 groups. Single crystals of 29 were grown from pentane at —30 °C and the structure was further confirmed by single—crystal X-ray crystallographic analysis. The molecular structure of 29 is shown in Figure 53. Selected bond distances and bond angles are given in Table 20. From the single X-ray structure of 29, Ir(l)-P(3),m,,, 10 3pm bond length is 2.359(2) A, Ir(1)-P(4),,,,., ,, 31w bond length is 2.334(2) A, and Ir(1)- 13(2),","510 H bond length is 2.319(2) A. The Ir-P bond length is directly proportional to the magnitude of the trans influence of the trans ligand; namely, the longer the Ir-P bond, the stronger the trans influence. It obviously shows that the order of trans influence is BPin> SiEt3 > H. To our knowledge, this is the first structural characterization of a transition metal boryl complex containing a silyl group. 90 Figure 53. ORTEP of fac-(PMe3)3Ir(H)(BPin)(SiEt3) (29). Thermal ellipsoids are shown at 25% probability. Table 20. Selected Bond lengths [A] and angles [°] for 29. Bond Distance [A] Bonds Angle [°] Ir(1)-B(1) 2.077(8) B(1)-Ir(1)-P(2) 92.5(2) Ir(1)-P(2) 2.319(2) B(1)-Ir(1)-P(3) 166.5(2) Ir(1)-P(3) 2.359(2) B(1)-Ir(1)-P(4) 82.9(2) Ir(l)-P(4) 2.334(2) B(1)-Ir(1)-Si(5) 85.1(2) Ir(1)-Si(5) 2.422(2) P(2)-Ir(1)-P(3) 100.9(7) P(2)-Ir(1)-P(4) 102.6(8) P(2)-Ir(1)-Si(5) 94.5(7) P(3)-Ir(1)-Si(5) 91.9(7) P(4)-Ir(1)-P(3) 95.9(7) P(4)—Ii(1)-Si(51 159.4(7) 91 Several novel iridium boryl complexes have been synthesized, and a strong trans influence of the boryl ligand has been observed in those complexes by comparing lr-P bond distances to the Ir-Pmm, to 3pm bond distance. In those complexes, significantly longer Ir-Pmm, 10 3pm bonds were observed compared to other Ir-P bonds in the complex. Ir-Pmm, to 3pm bond distances in those complexes are compared and summarized in Table 21 with the angles of PinB-Ir-PMe3(trans). Table 21. Comparisons of X2B-Ir-PMe3,,a,., to 3pm bond distances of iridium boryl complexes. Complex Bond Bond Length (A) Trans Ligands Angle (°) 17 PinB-Ir-PMe3 2.392( 1) PinB-Ir-PMe3 164.3(1) 18 PinB-Ir-PMe3 2.334(2) PinB-Ir-PMe; 178.0(2) 24 CatB-Ir-PMe; 2.399(1) CatB-Ir-PMe3 172.2(2) PinB—Ir-PMe3 2.35 2(4) PinB-Ir-PMe; 169.2(5) 25 PinB-Ir-PMe; 2.353(4) PinB-Ir-PMe3 169.0(5) PinB-Ir-PMe3 2.347(4) PinB-Ir-PMe3 169.6(5) 29 PinB-Ir-PMe3 2.359(2) PinB-Ir—PMe3 166.5(2) 92 Oxidation Chemistry of Ir1 complex with Boranes In contrast to the catecholboryl complexes, the corresponding compounds with other substituents (e. g. N, S) on boron are rather rare. In the past, B-H oxidative addition was exploited as a general synthetic method for preparing boryl compounds. A common theme in these reactions is the use of low-valent metal precursors containing readily dissociable ligands. This creates vacant sites in the metal coordination sphere, which allows for the oxidative addition of the boron reagents to the electron-rich metal center. This method was exploited to examine the reaction of several nitrogen-containing boranes including H[B(NH)2C(,H4], H[B(NH)2C10H6] (HBDAN), and H[B(NMe)2C6H4] with (PMe3)3Ir(COE)(Cl) (30). The results are summarized in Figure 54. Similar to the reactions of compound 30 with HBPin and HBCat, H[B(NH)2C6H4] reacts with compound 30 to give mer-(PMe3)3Ir[B(NH)2C61-I4](H)(Cl) (31) with the boryl group trans to Cl and H trans to PMe3 as the only observed product. Since, as discussed earlier, the borane approaches the metal center as an electrophile, H-B addition is preferred by bending Cl-Ir-P axis to form compound 31. Therefore, compound 31 is the kinetic product of the reaction. Consideration of the trans influence for the various ligands in compound 31 suggests that it is also the thermodynamic product. Changing the borane source to HBDAN gives two isomers mer-(PMe3)3Ir(BDAN)(H)(Cl) (32) and mer- (PMe3)3Ir(H)(BDAN)(Cl) (33) in a 90.4296 ratio. The geometries of these two complexes were determined by 1H, 1‘B, and 31P{‘H} NMR spectroscopy. The formation of the other isomer is presumably due to the bulkiness of the BDAN group, which might have interactions with methyl groups of a PMe3 ligand. It is expected that replacing the H 93 atoms on the N atoms of -[B(NH)2C6H4] with methyl groups would increase the steric bulkiness of the boryl group. We, therefore, examined the reaction of (PMe3)3Ir(COE)(Cl) (30) with H[B(NMe)2C6H4]. This reaction produced an isomer mixture of mer—(PMe3)3Ir[B(NMe)2C6H4](H)(Cl) (34) and mer- (PMe3)3Ir(H)[B(NMe)2C6H4](Cl) (35) in a ratio of l4.4:85.6. The geometries of these two complexes were determined by ]H, 1‘B, and 31P{'H} NMR spectroscopy. Characterization details of compounds 31, 32, 33, 34, and 35 are included in the experimental section. The reaction between compound 30 and H[B(NMe)2C6H4] suggests that steric factors can influence the outcome of boryl complex formation. (PMe3)4Ir(BPin) and fac-(PMe3)3Ir(BPin)3 have been prepared in order to examine their viability to be the OH activating species in the catalytic borylation reactions. Furthermore, Several novel Ir boryl complexes containing alkyl, aryl, or silyl ligands were synthesized and fully characterized. 94 H N f: H-B, D HN. ,NH N B H M P |I‘J134M93 PMe Ir COE Cl > e —r'- ( 3)3 ( X ) 25 °C, overnight M3 p" 93 CI (30) (31) ’1 CO 1. O H—B‘ H N N H H til 0 ‘3’ H PM K: H I ‘\PM63 I .‘ e3, (PM63)3IF(COE)(CI) o . = M63P71r.‘H + Me3P71r—B\ (30) 25 C, ovemrght Me3P Cl Me3P Cl '1 O (32) (33) 90429.6 Me ”I; § 1 ”'5‘ MeN NMe ‘ ’ H Me iie '13 ,PMea .PMealN (PMe3)3lr(COE)(CI) 25°C _ht = Me3P71r'——H + Me3P71r—B\ (30) '°"e""9 Me3P cr Me3P or Me (34) (35) 14.4:856 Figure 54. The reactions of (PMe3)3Ir(COE)(Cl) (30) with nitrogen-containing boranes including H[B(NH)2C6H4], H[B(NH)2C10H6] (HBDAN), and H[B(NMe)2C6H4]. 95 CHAPTER 5 PRELIMINARY MECHANISTIC STUDIES OF THE IRIDIUM/PHOSPHINE CATALYST SYSTEM FOR AROMATIC BORYLATION For many systems, catalysis by organometallic compounds is known to go through several of the following reactions: coordination of ligands to metals, oxidative addition, insertion and reductive elimination. In order to elucidate the mechanism of a homogeneously catalyzed reaction, it is a good idea to study the mechanism of each individual step in a series of relatively elementary chemical reactions by using tools such as kinetics, stereochemical studies, and spectroscopy. Irnportantly, each step must be shown to be kinetically and thermodynamically reasonable. Preliminary mechanistic studies of the Ir-catalyzed aromatic borylation were carried out and discussed in the previous chapter. Initially, it was deemed of importance to determine the viability of Ir1 and Ir'" boryl intermediates as the CH activating species in the Irm" and/or Ir"W cycles, respectively. Thus, the stoichiometric reactions of compounds 18nd 25 with benzene were examined. Stoichiometric Reactions of Ir1 and Ir Boryl Complexes with Arenes Thermolysis of (PMe3)4Ir(BPin) (18) was carried out in C6D6 at 150 0C, the reaction proceeded smoothly to give the corresponding iridium deuteride complex, (PMe3)4Ir(D), and C6D5BPin (Figure 55). 96 BFHn I \PMe3 _l ‘PMea ' M83P_!r'PMe3 C606 150°C M93P ir'PMezg * C5058P|n PMe3 PMe3 (13) Figure 55. Thermolysis of 18 in C6D6 at 150 °C. From 1H NMR spectra, more than 95% (PMe3)4Ir(D) was generated from stoichiometric reaction between 18 and C6D6 as shown in Figures 56 and 57. 97 b.o 0.0 m.o 0.." H!" «In m..." win min win h.d min Pb .lH—rr_- _—» _—_»___P__~Frbbh_hbebhphL_» P—pppr»_h_~_ 325.32% Eangezev .m_mbo§ofi 20.39 82.58% m2 Z I. ”Do of S eDeU E a“ mo max—9522. 6m gnaw:— 98 5.55 5532.: flour—BE 8: 803 Ema co; Hm.“ can Ema Nv._-mm._ 9:58 3on £26 .mmmboctofi Baa 83.50on MEZ I. ”Up of E eQeU E w— mo max—05.83% Km 2:”?— 99 Kinetic studies of the thermolysis of 18 in C6D6 were carried out and the reaction rate of the thermolysis was found to follow first-order kinetics as shown in Figure 58. 0 -0.5 l 5'0 1:: -1 § E" -1.5 y = -0.0416x - 0.0978 -2 R2 = 0.9962 -2.5 0 10 20 30 40 50 60 time (min) Figure 58. Plot of ln([18]t/[18]o) vs. time (min) for the thermolysis of 18 in C6D6 at 130 °C. There are two potential pathways to account for the stoichiometric transformation between 18 and C6D6 (Figure 59). One of them involves dissociation of PMe3 to generate a 16-electron intermediate, [(PMe3)3Ir(BPin)], followed by reaction with C6D6 to give products. The other potential pathway involves dissociation of two PMe3 ligands and generation of a 14-electron intermediate, [(PMe3)2Ir(BPin)], which subsequently reacts with C6D6 to give the final products. 100 k1 13 [(PMe3)3lr(BPin) + PMe3] (PMe3)2lr(BPin) + PMe3 1 16 e‘ K2 14 e‘ [(3 0606 k4 C606 (PMe3)4|r(D) + CngBPin (PMe3)4lr(D) + CngBPin Figure 59. Two potential pathways to account for the stoichiometric reaction between 18 and C6D6- From the steady-state approximation, the rate law of the reaction is derived as below. d[18] "T = Km [18] (6) k0 k1k-2kalceDellpMeal + k1k4[(313'-'313](k2 + kalceDeD (7) 05 = k-1k-2IPM9312 + (k-1k4 + szsllceoellpMeal + k41C606](k2 + kslceoel) Phosphine inhibition experiments were carried out to determine whether the reaction goes through a 16-electron intermediate or a 14-electron intermediate. Experiments of thermolysis of 18 in C6D6 in the presence of various concentrations of [PMe3] were examined. From 1/k0bs vs. [PMe3] plot (Figure 60), l/kobs obviously shows first order dependence on [PMe3]; therefore, the experimental data are consistent with the mechanism involving a 16-electron intermediate. 101 500 450 400 - 350 - 300 250 200 ,- 1 /kobs y = 3275.2x + 32.689 R2 = 0.9956 150 100 50 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 [PMes] (M) Figure 60. Plot of 1/k01,S vs. [PMe3] of the thermolysis of 18 in C6D6 in the presence of various concentrations of PMe3. From the results of phosphine inhibition experiments, a mechanism of the stoichiometric transformation is proposed in Figure 61. In the proposed mechanism, the reaction goes through phosphine dissociation to generate a l6-electron intermediate, [(PMe3)3Ir(BPin)], which can activate the CD bond of C6D6 to form [(PMe3)3Ir(BPin)(C6D5)(D)]. This intermediate reductively eliminate C6DsBPin to generate [(PMe3)3Ir(D)], which is followed by re-coordination of PMe3 to form the observed product, (PMe3)4Ir(D). 102 PMe3 - PMea BPin C D MesP Irvaes Me3P- 1r‘-PMe3 6 6 °PMe M p PMe3 3 e3 (13) M p ”1’88 6 r- M33P7' 6 5 M83 PMe -PMe3 D Me3P— ir.PMe3 —- Me3P- Ir- PMe3 PMe3 3 ’PM93 MeaP CeD5BPin Figure 61. Our proposed mechanism for the thermolysis of 18 in C6D6 Storchiometric reaction of 25 with arenes was investigated by the thermolysis of 25 in C6D6 at 150 °C. The thermolysis of 25 gave fac—(PMe3)3Ir(D)3 as the final iridium containing product and generated 3 equivalents of C6D5BPin. The thermolysis of 25 proceeded in a stepwise procedure as shown in Figure 62. BPin BP' I BPin * C6'36 l D.n Me3P;Ir- BPin —'__.’ Me3P- lr-BPin M93P PlMe3 - C6058P'n Me3P' PM63 1251 mm + CSDS ' CngBPin D BPin ID + C606 D “£1932; Ir D W Me3P; -|:r--D ' In 83 PMe3 6 5 MegPI PMe3 (36i-d23 Figure 62. The process of thermolysis of 25 in C6D6 at 150 °C. 103 The process of thermolysis of 25 was monitored by 1H, HB, and 3'P{]H} NMR spectra. Figure 63 displays the course of the reaction (the relative concentration of each species to C6Me6 (as an internal standard) versus time (in minutes)) as monitored by 1H NMR. 104 3 . faC-P3II‘B3 X . faC-P3II'B2D X A fac-P3IrBD2 x 2 5 ° fac-P3IrD3 . X C6D5BPin x o e; 2 *" <1) <3 2.. E e g '2 <6 8 “g 1.5 4 x 8 .9 m m a 9 8 9 a: .E 1 0 . x o - . o x 0.5 ° A A ’ x I - . A . A i A A ’ 0 I e e c k 3 + f a a 0 1500 3000 4500 6000 7500 Time (min) Figure 63. Concentration of each species relative to internal standard (C6Me6) vs. time (min) for the thermolysis of 25 in C6D6 at 150 °C measured by lH NMR. 105 The results showed that there was an induction period in the beginning of the thermolysis. Complex 25 was gradually converted to fac-(PMe3)3Ir(BPin)2(D) (21-d1) and generated 1 equiv. of C6D5BPin in the first step. Then fac-(PMe3)3Ir(BPin)2(D) (21-(11) was converted to fac-(PMe3)3Ir(BPin)(D)2 (36—d2) and the reaction generated a second equiv. of C6D5BPin. Finally, fac—(PMe3)3Ir(BPin)(D)2 (36-d2) was transformed to fac- (PMe3)3Ir(D)3 as the final iridium containing product and the reaction generated the third equiv. of C6D5BPin. The kinetic data for the thermolysis of 25 in C6D6 showed that the reaction did not follow first order kinetics (Figure 64). Furthermore, the thermolysis was strongly inhibited by external phosphine ligands. The inhibition phenomena are consistent with the observations in catalytic reactions where the reaction rate decreases dramatically when [P]:[Ir] ratio equals or exceeds 3:1. Detailed discussions will be included in the section followed by next paragraph. 106 I -O.5 — I In" 9'. :c: -1.5 e -2 _ I '2.5 I I F I I O 500 1000 1500 2000 2500 3000 Time (min) Figure 64. Plot of ln([25]1/[25]0) vs. time (min) for the thermolysis of 25 in C6D6 at 150 °C. 107 fac-(PMe3)3Ir(BPin)(H)2 (36) and fac-(PMeg)3Ir(BPin)2(H) (21) were prepared independently from the reaction of (PMe3)4Ir(H) (37) with HBPin and B2Pin2, respectively (Figure 65). Compound 37 reacted with HBPin at room temperature and generated a mixture of mer,cis-(PMeg)3lr(BPin)(H)2 (38) and fac-(PMe3)3Ir(BPin)(H)2 (36). After 6 days at room temperature, compound 36 was the predominant species in the reaction mixture. Similarly, compound 37 reacted with B2Pin2 at 60 °C and yielded a mixture of mer,trans-(PMe3)3Ir(BPin)2(H) (20) and fac—(PMe3)31r(BPin)2(H) (21). Compound 21 was the predominant species after the temperature was increased to 100 0C for 7 hours. The geometries of 20, 21, 36, and 38 were determined by 1H, HB, and 3 IP {'H} NMR spectroscopy and summarized in the experimental section. -PMe3 '1' PMe3 F.) M33 (PMe3)4lr(H)+ HBPin 25°C MeaP;111-1 .— Me3P;Ir‘-H (37) M8313 BPin 25 C M6313 éPin (33) (36) BPin . 'PMea l.PMes 7533111 (PM93)1|r(H)+ B2P|nz W Me3P;llr’-H 100°C 711 Me3P-(—1’r‘—BPin (37) MesP BPin ' Me3P H (20) (21) Figure 65. The reaction of (PMe3)4Ir(H) (37) with HBPin and B2Pin2. Correlation between Phosphine Ligands and Catalytic Activity Since both IrI and Ir"1 boryl complexes can effect stoichiometric borylation of benzene, they are potentially viable candidates to be the C-H activating species for 108 catalytic borylation reactions. However, the stoichiometric reactions of (PMe3)4Ir(BPin) (18) and fac-(PMe3)3Ir(BPin)3 (25) with benzene exhibit dramatically different reactivities in terms of phosphine dependence. The thermolysis of Ir!" complex 25 in C6D6 is strongly inhibited by external phosphine ligands, whereas the thermolysis of Ir1 complex 25 shows inverse first-order dependence on [PMe3]. For catalysis to occur it is usually required to produce vacant coordination sites in organometallic complexes. Since both compounds are coordinatively saturated 18-electron complexes, it is not surprising that the stoichiometric reaction of 18 or 25 with benzene proceeds slowly at elevated temperature. As previously mentioned it was shown that a phosphine dissociation pathway is responsible for generating active species to activate C-H bonds of benzene in the stoichiometric reaction of 18 with benzene. Furthermore, in the catalytic borylation reactions, the reaction rate was shown to be dependent on the nature of phosphine ligands and the relative concentration between phosphine ligands and iridium metal complex.57 In particular, borylation rates were appreciable when [P]:[Ir] < 3:1 but decreased dramatically when [P]:[Ir] ratio equaled or exceeded 3:1. The fact that phosphine inhibition is observed in the thermolysis of complex 25, which contains three PMe; ligands, in the presence of external phosphine ligands is consistent with the inhibition phenomena seen in catalytic reactions when [P]:[Ir] ratio equals or exceeds 3:1. The thermolysis of IrI complex 18, however, shows inverse first order dependence on [PMe3]. The reaction rate decreases when the concentration of PMe3 increases. Since the thermolysis of 18 is not completely shut down when [P]:[Ir] >> 3, the observation is not not consistent with that in catalytic reactions. These experimental observations support a 109 mechanism involving an Ir!“ boryl intermediate being the C-H activating species in these catalytic borylation reactions. Substantially improved catalytic activity was observed with the use of bidentate chelating phosphine ligands including 1,2-bis(diphenylphosphino)ethane (dppe) and 1,2- bis(dimethylphosphino)ethane (dmpe). This observation strongly supports the viability of bisphosphine intermediates. A further piece of evidence is that the 18-electron bisphosphine compound, (PMe3)2Ir(H)5, is an effective pre-catalyst for aromatic lll borylation. Those data imply a mechanism involving Ir and Irv intermediates in a Ir"W catalytic cycle. Stoichiometric and Catalytic Borylations of Iodobenzene Previously it was found that iridium catalysts generated from an Irl source, (Ind)Ir(COD) (13) and dppe, were ineffective for the aromatic borylation of iodobenzene. However, iodobenzene and HBPin reacted smoothly to yield an isomer mixture of C6H4(I)(BPin) when active catalysts were generated from an Irm source, (MesH)Ir(BPin)3 (14) and dppe. Both the 11’ boryl complex, (PMe3)4Ir(BPin) (18), and the 11‘" boryl complex, fac-(PMe3)3Ir(BPin)3 (25) can effect stoichiometric reactions with benzene to produce PhBPin and the corresponding hydride complexes. However, the arene products from stoichiometric reactions of 18 and 25 with iodobenzene differ substantially. Specifically, compound 18 reacted rapidly with iodobenzene at room temperature to form an off-white precipitate, but isomers of C6H4(I)(BPin) were not detected, even after prolonged thermolysis. Analysis of the 1H and 31P{1H} NMR spectra of the off-white 110 precipitate from the reaction indicated that the material contained a number of species, which could not be identified (Figure 66). Presumably, the reaction proceeded via oxidative addition of iodobenzene to complex 18 to form [(PMe3)3Ir(I)(C6H5)(BPin)] followed by several potential reaction pathways to generate a variety of products. For example, [(PMe3)3Ir(I)(C6H5)(BPin)] could eliminate C-B or I-B bonds. 1H NMR spectrum ': '. ,/i “ / _. l ' . r‘ \ ‘5 ”A- __ _ «‘._~ ‘ _, 14‘1"} L]. Jr»): ‘_A 1K, J‘\_ 1.. .44 ‘~ . . '_,/ ‘v‘ i r ‘ \1 ‘ l 2.60 2.40 2.20" “”2100“ '1."80'T‘1.60 ' 1.40” ”1.20 ‘ 1'B NMR spectrum 1 1 "78"”63“ 5040 "30 2‘0 10’ '0 8 11’0" 1208180140 £63160" 470 31P{1H} NMR spectrum Figure 66. 'H, “B, and 3‘P{‘H} NMR spectra of the off-white precipitate from the reaction between 18 and C6H5I. lll Conversely, thermolysis of 25 in iodobenzene at 150 oC produced, in addition to a 45% yield of PhBPin, m- and p-C6Ha(I)(BPin) in 54% yield (Figure 67). The significant quantities of PhBPin formed were presumably generated from a competitive C-I activation pathway followed by PhBPin reductive elimination. fac-(PMe3)3Ir(BPin)3 150°C 4* \—\/ I + Q—BPin (25) BPin (54% 60 yield) (45% 60 yield) Figure 67. Thermolysis of 25 in iodobenzene. The previous paragraph demonstrates that borylation products of iodobenzene are not obtained when IrI sources are used under stoichiometric and catalytic conditions, whereas, Ir"l complexes effect both stoichiometric and catalytic borylations. Furthermore, in situ generation of “Ir”I species” from an IrI source, compound 13, and dppe has been demonstrated to be a viable way to generate effective catalysts for borylation of iodobenzene. These experimental observations are consistent with a mechanism 111 involving Ir and IrV intermediates. Kinetic Isotope Effects in Aromatic Borylation 112 The isotope effects involved in the activation of arene C-H bonds by the intermediate [Cp*Rh(PMe3)] have been investigated by Jones and F eher.5 8 In their study of reductive elimination of arenes from aryl hydride complexes Cp*Rh(PMe3)(Aryl)(H), they discovered that a low—energy pathway existed for the interconversion of the carbon attached to the metal. The isomerization was found to occur in a sequential [1,2] fashion, and it can be accommodated by the reversible formation of an nz-arene complex as shown in Figure 68. i9: ‘59:: 592: I /Rh"'PMe3 —— Rh _. H,Rh-"PMe3 H I "'PMea { > 1) Figure 68. The process of reversible formation of nz-arene complexes proposed by Jones and Feher. The photolysis of Cp*Rh(PMe3)(H)2 offers a convenient method for the photoinduced generation of the coordinatively unsaturated intermediate [Cp*Rh(PMe3)] by elimination of dihydrogen. The active species is able to activate OH or CD bonds under conditions of kinetic control. Jones et al. conducted a kinetic isotope experiment by irradiating Cp*Rh(PMe3)(H)2 in a 1:1 (vzv) mixture of CsHs/C6D6 at 10 0C, which resulted in the evolution of H2 and the formation of the benzene C-H/C-D activation products Cp*Rh(PMe3)(C6H5)(H) (4) and Cp*Rh(PMe3)(C6D5)(D). Quenching of the reaction with CCla forms the corresponding chloro derivatives Cp*Rh(PMe3)(C6H5)(Cl) 113 and Cp*Rh(PMe3)(C6D5)(Cl). Mass spectral analysis of this mixture revealed a 1.05:1 ratio of the products in which C6H6 versus C6D6 has been activated (Figure 69). Since benzene is not labile in the products Cp*Rh(PMe3)(C6Hs)(H) and Cp*Rh(PMe3)(C6D5)(D) under the condition of the experiment (10 °C), this ratio reflects the kinetic isotope effect for arene complexation and/or C-H bond activation. This small value of kszD was consistent with there being little or no C-H bond breaking in the rate- determining step. kHIkD = 1.05 Z 1.00 Figure 69. Observed kH/kD in the activation of a 1:1 mixture of Cng/CbDb by the intermediate [Cp*Rh(PMe3)]. In order to access the C-H bond-breaking step, Jones et a1. carried out a similar kinetic isotope effect experiment with 1,3,5-C6D3H3 and kH/kD was determined to be 1.4. Because complexation of [Cp*Rh(PMe3)] to 1,3,5-C6D3H3 can produce one possible 112- arene complex, the ratio kH/kD for this reaction reflects only the isotope effect involved in the cleavage of the C-H. Since both experiments involve bimolecular activation of the benzene C-H/C-D bond ([Cp*Rh(PMe3)] + benzene), their observation of different kinetic isotope effects for these two experiments prove that a direct insertion of [Cp*Rh(PMe3)] into the OH bond of benzene is not occurring. In other words, they have 114 determined that the activation of arene C-H bonds by [Cp*Rh(PMe3)] proceeds through an intermediate complex, [Cp*Rh(PMe3)(n2-C(,I{6)]. In order to have a deeper understanding of the mechanism of borylation by (Ind)Ir(COD) (13)/2 PMe3 or (MesH)Ir(BPin)3 (14)/2 PMe3 pre-catalyst system and hopefully identify the rate-determining step in the catalytic borylation reactions, borylation reactions in a molar ratio 1:1 mixture of C6H6/C6D6 and separately in 1,3,5- C6D3H3 were carried out. If a kinetic isotope effect is observed for the borylation reaction in a molar ratio 1:1 mixture of C6H6/C6D6, it may result from coordination of C6H6 vs. C6D6 or from the C-H/C-D bond breaking step. On the other hand, the kinetic isotope effect for the borylation reaction of 1,3,5-C6D3H3 can only result from the C-H bond activation step since there is only one arene for coordination. The results from experiments of catalytic reactions are summarized in Table 22. 115 Table 22. Borylation reactions with HBPin in a molar ratio 1:1 mixture of C6H6/C6D6 or 1,3,5-C6D3H3 catalyzed by Irl and IrIll sources at 150 °C, [Ir] = 2 mol%, [PMeflzLIr] = 2: 1. Entry Pre—catalysts Substrate Product Distribution” (MesH)Ir(BPin)3 (14) Molar ratio 1:1 CstBPinszHsBPin 1 rr'" (gm/C11)6 1.00:2.28 (100% conversion) 14 C6D2H3(BPin):C6D3H2(BPin) 2 193,59'C6D3H3 Ir'" 1001.94 (100% conversion) (Ind)Ir(COD) (13) Molar ratio 1:1 CgDsBPinzCoHsBPin 3 Ir' c.5H1,/(:6D6 1002.29 (100% conversion) 13 C6D2H3(BPin):C6D3H2(BPin) 4 1,3,5,-C6D3H3 rr' 1.00:2.06 (100% conversion) Comparison of the results in Table 22 shows similar kH/kD values for the borylations of C6H6/C6D6 (1:1) and of 1,3,5-C6D3H3 by Irl or Irm. Furthermore, it appears that the kinetic isotope effect is essentially identical from the IrI and Ir"1 pre-catalyst sources, which suggests a similar or same active species. The large observed kinetic isotope effect in borylations of CeHe/CeDs (1:1), which is different from Jones’ case, suggests that arene coordination cannot be the rate-determining step. If the arene coordination is the rate—determing step, we expect to see no or, at most, a very small kinetic isotope effect since it does not involve the OH bond breaking event. However, current data cannot distinguish whether C-H bond breaking or reductive elimination of the arylboronic ester is the rate-determining step (see Figure 73). If reductive elimination 116 of the arylboronic ester were the rate-determining step, we would expect to see a small secondary isotope effect because the reaction does not involve cleavage of C-H/C-D bonds. The expected small isotope effect differs from relatively large observed kH/kD (2.06 and 1.94) for the borylations in 1,3,5-C6D3H3 fi'om Irl and Irl" sources, respectively. However, it is important to remember that any step prior to the rate-determining step can contribute to the observed kinetic isotope effect. Therefore, the current data cannot definitely determine the actual rate-determining step in the borylation reactions. Stoichiometric borylations of (PMe3)4Ir(BPin) (18) and fac—(PMe3)3Ir(BPin)3 (25) in a 1:1 mixture of CfiHb/CfiDfi and separately in 1,3,5-C6D3H3 were also examined. Thermolysis of 25 in a 1:1 mixture of C6H6/C6D6 and separately in 1,3,5-C6D3H3 at 150 oC gave similar kinetic isotopes effect kH/kD as compared to the catalytic borylation reactions, whereas thermolysis of 18 in a 1:1 mixture of CgHg/C6D6 and separately in 1,3,5-C6D3H3 at 150 °C gave relatively larger kinetic isotope effect kH/kD than those observed in catalysis as shown in Table 23. 117 Table 23. Stoichiometric borylation reactions of 18 and 25 with a molar ratio 1:1 mixture Ofchb/C6D6 OI‘ 1,3,5-C6D3H3 at 150 °C. Entry Iridium Complex Substrate Product Distribution59 fac-(PMe3)3Ir(BPin)3 Molar ratio 1:1 C6DsBPin:C6H5BPin l (25) C6H6/C6D6 1.00:2.53 (100% conversion) 11,111 25 C6D2H3(BPin) : C6D3H2(BPin) 2 193959-C6D3H3 Ir'” 1001.93 (100% conversion) (PMe3)4Ir(BPin) (18) Molar ratio 1:1 C6D5BPin:C6H5BPin 3 Ir] CdidC6D6 1.00:2.67 (100% conversion) 18 C6D2H3(BP11’1)2C6D3H2(BP111) 4 1,3,5,-C6D3H3 Ir' 1002.37 (100% conversion) Mechanistic Discussions The existence of an arene coordination step in the catalytic cycle of aromatic borylation by 13/2 PMe3 or 14/2 PMe3 pre-catalyst systems cannot be completely ruled out based on the present data. However, Miyaura26f and co-workers reported the isolation of a trisboryl complex [Ir((dtbpy)(COE)(BPin)3] as shown in Figure 70, which is chemically and kinetically competent to be an intermediate in the catalytic process in their pre-catalyst system (3 mol% 1/2[IrCl(COE)2]2/dtbpy). Based on a lesson from 118 Halpem’s work in elucidating the mechanism of the hydrogenation reaction catalyzed by Wilkinson’s catalyst, (PPh3)3Rh(Cl), isolation of a stable olefin complex is inconsistent with an 112—arene complex pathway. Therefore, the arene coordination step prior to C-H bond activation seems unlikely. [ll’(COE)2Cl]2 * + szlnz mesitylene Figure 70. A trisboryl complex [Ir((dtbpy)(COE)(BPin)3] isolated by Miyaura and co- workers. There are two potential pathways for the active species to react with an arene. One possibility is that it proceeds through oxidative addition of a C-H bond of an arene to give an Irv intermediate followed by reductive elimination of an arylboronate ester to give the corresponding hydride complex. Another option is that the corresponding hydride complex is formed in a one-step “o-bond metathesis” reaction as shown in Figure 71. o-bond metathesis has been confirmed only for (10 complexes of the early transition metal and lanthanides, where oxidative addition is precluded. 119 PinB \BPin (PR312‘-‘- + Ph- 4"\ Ph - Ph-BPin PinB H 111 . IrV intermediate [(PR3)2lr (BPIn)3l C-H activating species +m 1 ,BPin‘ 1/Ph-:3Pin [(PRsizlr“'(BPin>21Hil (PR3)2(BPin)2'r~/.‘ ’:Ph \H’ o-bond metathesis transition state Figure 71. Possible mechanisms for Ir"I borylation reaction. Although we have not yet investigated the mechanism of the borylation extensively, we have obtained qualitative information regarding this question. From the kinetic studies of the stoichiometric reaction of (PMe3)4Ir(BPin) (18) with benzene, we established that the intermediate [(PMe3)3Ir(BPin)] activates the C—H bond of benzene. Furthermore, in the stoichiometric reaction of fac-(PMe3)3Ir(BPin)3 (25) with benzene, a transient species similar to [(PMe3)2Ir(BPin)3], presumably generated in the reaction mixture, activates the C-H bond of benzene. Therefore, oxidative addition of a C-H bond of an arene to an iridium boryl complex is a viable reaction pathway. In order to evaluate PhBPin reductive elimination, we examined the thermolysis of compound 28 in C6D6. Compound 28 has the BPin group trans to the phenyl group and it is stable at room temperature. Before reductive elimination of PhBPin occurs, compound 28 is expected to isomerize to another isomer with the BPin group and the phenyl group in a cis geometry. Thermolysis of compound 28 in C6D6 at 50 °C was carried out and the reaction was 120 monitored by 1H, 11B, and 31P{1H} NMR spectroscopy. At 50 OC, compound 28 was converted to a mixture of fac-(PMe3)3Ir(C6D5)(D)(H) (39), fac-(PMe3)3Ir(BPin)(H)2 (36), and (PMe3)4Ir(H) (37), and PhBPin was produced. After 55 hours, compound 28, 39, 36, and 37 were in the ratio of 16:73:11:<1 (Figure 72). BPin I APM83 C6061 " PhBPin Me3P711—H o = M6313 FLh 50 C, 55 h ,PMe3 Me3P;|r -H Me3P (28) 1 0506 (28)3(39)=(36)I(37) = 16:73:11:<1 Figure 72. Thermolysis of compound 28 in C6D6 at 50 °C. Presumably reductive elimination of PhBPin from compound 28 generates an intermediate, [(PMe3)3Ir(H)], which can subsequently activate a C-D bond of CbDr, to form compound 39. Compound 36 most likely comes from the oxidative addition of HBPin to the intermediate [(PMe3)3Ir(H)]. HBPin could come from a minor pathway involving H-B reductive elimination from compound 28. The generation of PhBPin from 121 the thermolysis of compound 28 in C6D6 shows that PhBPin reductive elimination is a viable pathway as well. In addition to some other observations discussed previously: (1) Borylation products of iodobenzene are not obtained when IrI sources are used under stoichiometric and catalytic conditions, whereas Ir"I complexes effect both stoichiometirc and catalytic borylations. (2) Improved catalytic activity is observed with chelating phosphines and inhibition is observed when [P]:[Ir] ratios equal or exceed 3:1, strongly supporting the viability of bisphosphine intermediates. (3) The l8-electron bisphosphine compound, (PMe3)2Ir(H)5, is an effective pre-catalyst for borylation. Therefore, we presently favor the simple scheme which involves a direct oxidative addition of a CH bond of an arene to the proposed active species, [(PR3)2Irm(BPin)3] to form an lrV intermediate, [(PR3)2Ir"'(BPin)3(H)(Ph)]. Reductive elimination of PhBPin from the Irv intermediate, [(PR3)2Ir"‘(BPin)3(H)(Ph)], gives [(PR3)2Irm(BPin)2(H)], which converts to [(PR3)2IrV(BPin)3(H)2] in the presence of HBPin. After releasing H2, the reaction regenerates [(PR3)2Irm(BPin)3], the proposed C-H activating species, to complete the catalytic cycle as shown in Figure 73. 122 H2 [(PR3)2|r"'(BPin)3] ch6 C-H activating species HBPin PinB OBPin (PR31211"'(H)18P1n>2 (P8312 11‘“ . /1 PinB H QBPin Figure 73. A putative mechanism for aromatic borylations catalyzed by iridium boryl complexes. An interesting reactivity was observed in the thermolysis of fac-(PMe3)3lr(BPin) 3 (25) in C6D6 in the presence of 10% of (MesH)Ir(BPin)3 (14). The reaction proceeded at 100 °C instead of occurring at 150 °C to generate a mixture of fac-(PMe3)3Ir(BPin)2(H/D) and fac-(PMe3)3Ir(BPin)(H)(D) in approximately 62:38 ratio and C6D5BPin after 33.5 hours. The hydride presumably comes from a PMe3 ligand. As has been discussed earlier, the mesitylene ligand of compound 14 is very labile presumably due to the very strong trans influence of the BPin group, therefore, it is most likely that 14 acts as a phosphine trap, which facilitates the PMe3 dissociation from compound 25 in the thermolysis 123 reaction of 25 in C6D6. The actual role of complex 14 in the reaction and the difference in terms of reactivity require further elucidation. Miyaura, Ishiyama, and co-workers60 recently developed a new catalyst system (1.5 mol% [Ir(OMe)(COD)]2/ 3 mol% dtbpy) for borylation of arenes at room temperature with a stoichiometric amount of boron reagent (B2Pin2) and arene to produce the corresponding arylboronates in high yield. This new system has extended the functional group tolerance to a CN group. The high catalyst efficiency of [Ir(OMe)(COD)]z can be attributed to the more facile formation of iridium boryl complexes. Iridium triboryl complexes have been implied as possible intermediates in the borylation of arenes. Presumably, the transmetallation reaction between B2Pin2 and [Ir(OMe)(COD)]z results in the formation of an Irl boryl complex after releasing MeOBPin. The Irl boryl complex then undergoes oxidative addition of B2Pin2 to yield an Ir'" triboryl complex as shown in Figure 74. + szlnz + szinz gpin I -OR + B P' | -BP' ——> — ' [r] 2 1n2 _ MeOBPin [r] In [éréinBPln [Ir] = [lr(bpy)(COD or COE)] Figure 74. Iridium tris(boryl) intermediate in borylation reactions. The results from preliminary mechanistic studies on the new iridium catalyst [V . I“ catalytic system indicate that a mechanism involving Ir!" and Irv intermediates in an Ir cycle is most likely. Correlations between the stoichiometric and catalytic reactions provided a deeper insight into the mechanism of aromatic borylation. 124 Our studies allow arylboron species to be prepared in an economical fashion and also demonstrate that C-H bond activation and functionalization can be developed into a practical synthetic tool. Future work is needed to further elucidate the mechanism for aromatic borylations catalyzed by iridium boryl complexes in order to have a deeper understanding for this important transformation. With the rapid advances in this area, we expect that Ir-catalyzed borylations of aromatics will certainly find acceptance in the synthetic arsenal of organic chemists. 125 CHAPTER 6 EXPERIMENTAL General Considerations All manipulations were performed using glove box, Schlenk, or vacuum-line techniques. Pentane, diethyl ether, and tetrahydrofirran were pre-dried over CaCl2 and distilled from Sodium/benzophenone ketyl. Toluene and benzene were pre-dried over CaCl2 and distilled from Sodium metal. Methylene chloride was pre—dried over C aC l 2 and distilled from CaH2. Cyclohexane was purified and dried according to the method reported by Perrin and Armarego.6| Hexamethyldisiloxane, decane, and dodecane were distilled from sodium metal. CDCl3, tetrahydrofuran-d3, p-xylene-dm, cyclohexane-d12, and 1,3,5-C6D3H3 were dried with and vacuum transferred from 3A sieves. Benzene-d6 and toluene-d3 were dried with, vacuum transferred from 3A sieves, and stored over a sodium mirror. CD2C12 was dried with 4A sieves. HBPin was purchased from Aldrich, further purified by stin'ing with PPh; to remove BH3, and then vacuum distilled to give the borane as a clear viscous liquid. B2Pin2 was purchased from Frontier Scientific and used as received. PMe3 and PMe3-d9 were purchased fi'om Aldrich and vacuum transferred to an air—free flask respectively before use. PEt3, P'Pr3, P’Bug, PCy3, PPh3, 1,2-Bis(diphenylphosphino)ethane (dppe), 1,2- Bis(diphenylphosphino)propane (dppp), Bis(dimethylphosphino)methane (dmpm), 1,2- Bis(dimethylphosphino)ethane (dmpe), 1,2-Bis(dicyclohexylphosphino)ethane (dCype), 1,2-Bis(diphenylphosphino)benzene (dppb), 2,2’-dipyridyl (bpy), 4,4’-di-tert-butyl-2,2’- 126 dipyridyl (dtbpy), 2-(diphenylphosphino)-2’-(N,N-dimethylamino)biphenyl, 2- (dicyclohexylphosphino)-2’-(N,N-dimethylamino)biphenyl, 1,10-phenanthroline, and 2,2’-bithiophene were used as received from commercial sources. 1,2-dimethoxyethane, N,N,N,N-tetramethylethylenediamine (TMEDA), thiophene and triethylsilane were distilled prior to use. Substrate: anisole, N,N-dimethylaniline, 2,6-lutidine, benzotrifluoride, N,N- diethylbenzamide, ethylbenzoate, 1,3,5-C6H3F3, and 4-fluorobromobenzene were distilled and further dried by passing through a column of activated alumina prior to use. C6HF5 and l,3-bis(trifluoromethyl)benzene were distilled and then dried over 4A sieves, followed by vacuum transfer to an air-free flask. m-xylene, p-xylene, o-xylene, fluorobenzene, chlorobenzene, bromobenzene, 1,2-dichlorobenzene, 1 ,3- dichlorobenzene, 1,3-dibromobenzene, 4-chlorobenzotrifluoride, 4-fluoroanisole, 4- chloroanisole, 4-chlorotoluene, 2-chloroanisole, 2-chlorotoluene, and 2-methylanisole were distilled from Sodium metal. Veratrole and l,4-bistrifluoromethylbenzene were distilled from CaH2. 1,4—dichlorobenzene was sublimed prior to use. Starting materials: [Ir(COD)Cl]2,62 [Ir(COE)2Cl]2,62 (Ind)Ir(COD),“ (PMe3)41r(Cl),64 (PMe3)4Ir(H),44 (PMe3)4Ir(Me),52 (PMe3)3Ir(Ph),55 Cp*Rh(n4-C6Me6),65 Cp*Irtn4-C6Me61,“ Cp*1r(PMes)(H)2,67 Cp*IrtPtCDihxrlh,“ (CsMe4Et)Ir(PMes)(H)2,"7 (PMe3)2Ir(H)5,68 Cp"‘IrH4,69 and Cp"‘Ir(PMe3)(H)(BPin)21 were prepared according to literature methods. 1H and 13C{'H} NMR spectra were recorded on Varian Inova—300, VXR-SOO, or Inova—600 spectrometers and referenced to residual proton solvent signals. llB, 19F, and 31P{lH} spectra were recorded on Varian VXR—300 or Varian [nova—300 spectrometers, 127 operating at 96.29, 282.35, and 121.49 MHz respectively and referenced to external standards. Boron chemical shifts were referenced to a neat BF 3-Et2O external standard. Fluorine chemical shifts were referenced to a neat CFC13 external standard. Phosphorous chemical shifts were referenced to an 85% phosphoric acid external standard. Elemental analyses were performed at Michigan State University using a Perkin Elmer Series II CHNS/O Analyzer 2400. GC-MS data were obtained using a HP 61 800A GCD system. Syntheses (MesH)Ir(BPin)3 (l4). Complex 13 (2.54 g, 6.11 mmol) and HBPin (3.91 g, 30.55 mmol) were dissolved in 32 mL mesitylene. The light brown solution was then heated at 75 °C for 48 h. The reaction mixture turned to dark brown after 24 h at 75 °C. Mesitylene was removed under high vacuum to give viscous dark brown oil. The crude mixture was then triturated with hexamethyldisiloxane (3 x 2 mL). A white solid (797 mg, 19%) was obtained after filtration and washed with cold hexamethyldisiloxane. mp 140 °C (dec). IH NMR (600 MHz, Cst) 8 1.33 (s, 36H, 3 BO2C6H12), 2.24 (s, 9H, 3 CH3), 5.62 (s, 3H, 3 CH). ”B NMR (Cst) 6 32.5. '3C{‘H} NMR (300 MHz, C6D6) 6 19.7 (s), 25.7 (s), 81.0 (s), 97.0 (s), 118.05 (8). Anal. Calcd for C27H43kB206: C, 46.77; H, 6.98. Found: C, 47.13; H, 7.18. mer-(PMe3)3Ir(BPin)(H)(CI) (15). A solution of PMe3 (102 mg, 1.34 mmol) in 2 mL THF was added dropwise to a solution of [Ir(COE)2Cl]2 (200 mg, 0.22 mmol) in 4 mL THF solution. The reaction mixture was stirred at room temperature for 2 h. The solvent was then removed under reduced pressure. The residue was redissolved in 6mL 128 THF and a solution of HBPin (65 11L, 0.45 mmol) in 2 mL THF was added dropwise. The reaction mixture was stirred at room temperature overnight. Next day, the solution was filtered through celite to remove trace suspension. The filtrate was pumped down to give a spectroscopically pure white solid (243.0 mg, 93%). The product can be recrystallized from concentrated THF solution at —30 °C to give colorless crystals. mp 170-172 0C (dec). 1H NMR (C6D6) 5 -9.66 (dt, J= 136.4 Hz, 21.8 Hz, 1H, hydride), 8 1.04 (s, 12H, BO2C6H12), 1.37 (d, J = 7.9 Hz, 9H, PMe; trans to hydride), 1.62 (t, J = 3.7 Hz, 18H, 2 PMe3 trans to each other). l3C(‘H1 NMR (C6D6) 6 18.4 (d, J: 26.7 Hz), 20.7 (dt, J: 3.5 Hz, 19.1 Hz), 25.8 (s), 80.8 (s). ”B NMR (C6D6) 6 28.5. 3‘P1‘H} NMR (C613,) 6 —46.1 (t, J = 20.6 Hz, 1P, PMe3 trans to hydride), —40.4 (d, J = 21.4 Hz, 2P, 2 PMe3 trans to each other). Anal. Calcd for C15H4oIrBClO2P3: C, 30.86; H, 6.91. Found: C, 30.91; H, 7.06. mer,cis-(PMe3)3Ir(BPin)2(Cl) (17). A solution of B2Pin2 (191 mg, 0.75 mmol) in 5 mL THF was added to a suspension of (PMe3)4Ir(Cl) (400 mg, 0.75 mmol) in 30 mL THF in a schlenk tube. The reaction mixture was heated at 70 °C for 1 day. The orange suspension gradually changed to a gray color suspension. The reaction mixture was cooled down and filtered through celite to remove gray precipitates. The filtrate was then pumped down to give a colorless crystalline solid. The crude product was recrystallized from pentane at —30 °C. The product was collected as colorless crystals (390 mg, 73%). mp 156-158 °C (dec). 1H NMR (C6136) 6 1.10 (s, 12H, 302C6H12), 1.22 (s, 12H, BO2C6H12), 1.31 (d, J = 6.9 Hz, 9H, PMe-5 trans to BPin), 1.72 (t, J = 3.6 Hz, 18H, 2 PMe3 trans to each other). 13C(‘H} NMR (Cst) 6 15.9 (t, J: 19.1 Hz), 19.4 (d, J: 41.3 Hz), 25.4 (s), 81.9 (s). ”B NMR (C6D6) 6 28.0, 36.5. 3‘P1‘H} NMR (C6D6) 6 -514 (br, 129 1P, PMe3 trans to BPin), -41.1 (d, J = 26.9 Hz, 2P, 2 PMe3 trans to each other). Anal. Calcd for C21H51IrB2C104P3: C, 35.53; H, 7.24. Found: C, 35.48; H, 7.60. (PMe3)4lr(BPin) (18). A solution of PMe; (161 mg, 2.12 mmol) in 4 mL THF was added to a solution of 17 (500 mg, 0.7 mmol) in 5 mL THF in a schlenk tube. A solution of KO’Bu (158 mg, 1.4 mmol) in 5 mL THF was then added to the reaction mixture. The reaction mixture was stirred at room temperature for 90 min. The solvent was removed under reduced pressure. The product was extracted with 8 mL pentane. The pentane filtrate was then pumped down to give a white solid (402 mg, 92%). The material always contained a small amount of (PMe3)4Ir(H) (37) (ca. 3% by 1H NMR) due to its considerable moisture sensitivity). The product was recrystallized from a concentrated pentane solution at -30 0C to give colorless crystals. mp 130-137 °C (dec). 1H NMR (Cng, 25 °C) 6 1.24 (s, 12H, 302C6H12), 1.58 (br s, 36H, PMe3). ‘3C1'H1 NMR (C601,, 25 °C) 6 26.8 (s), 28.9 (m), 81.0 (s). ”B NMR (C6D6) 6 38. 3'P{'H} NMR (C6D6, 25 °C) 5 —57.5 (br s, 4P). Anal. Calcd for C13H431IBOZP4Z C, 34.67; H, 7.76. Found: C, 34.76; H, 7.89. fac—(PMe3)3Ir(BPin)3 (25). A solution of PMe3 (220 mg, 2.9 mmol) in 2 mL CgHr, was added a solution of 14 (400 mg, 0.58 mmol) in 4 mL C6116 in a vial. The reaction mixture was stirred at ambient temperature for 30 min and the solvent was removed under reduced pressure to give a white solid (461 mg, 99%). The product was recrystallized from a concentrated pentane solution at —30 °C to give colorless crystals. mp 184 °C (dec). 1H NMR (CeDe) 5 1.34 (s, 36H, BOzCoHiz), 1.52 (m, 27H, PMe3). l3C(‘H} NMR (C6D6) 23.7 (m), 26.5 (s), 80.4 (s). “B NMR (C6D6) 6 36.0. 3"1>{‘H} NMR 130 (C6D6) 8 —64 (br, 3P). Anal. Calcd for C27H63Ir8306P3: C, 40.47; H, 7.92. Found: C, 40.72; H, 8.01. fac-(PMe3)3Ir(BPin)(H)(Me) (26). A solution of HBPin (27 mg, 0.21 mmol) in 2 mL pentane was added to a solution of (PMe3)4Ir(Me) (100 mg, 0.21 mmol) in 4 mL pentane. The reaction mixture was stirred at ambient temperature for 5 min and the solvent was then removed under reduced pressure to give an isomer mixture of fac- (PMe3)3Ir(BPin)(H)(Me) (26) and mer-(PMe3)3Ir(Me)(H)(BPin) (27) in a ratio of 83:17 (94 mg, 75%). fac—(PMe3)3Ir(BPin)(H)(Me) (26). 1H NMR (CsDe) 8 -11.30 (dt, J = 140.4 Hz, 18.9 Hz, 1H, hydride), 0.40 (m, 3H, Me), 1.17 (d, J = 6.4 Hz, 9H, PMe3 trans to BPin), 1.25 (s, 12H, BOzCsHiz), 1.35 (d, J = 7.3 Hz, 9H, PMe3 trans to hydride), 1.47 (d, J = 7.9 Hz, 9H, PMe3 trans to Me). ‘3C{‘H} NMR (CgDo) 6 -36.3 (dt, J = 62.5 Hz, 7.6 Hz, Me), 19.9 (ddd, J= 24.7 Hz, 5.5 Hz, 2.7 Hz), 20.9 (dt, J= 21.3 Hz, 2.7 Hz), 22.0 (td, J = 18.5 Hz, 4.1 Hz), 25.8 (s), 25.9 (s), 80.1 (s), 80.2 (s). 1]B NMR (C6D6) 8 38.6. 3'P{'H} NMR (C6D6) 6 -63.3 (br, 1P, PMe3 trans to BPin), -56.83 (dd, J = 13.4 Hz, 23.2 Hz, 1P, PMe3 trans to hydride), -55.16 (dd, J = 13.4 Hz, 18.3 Hz, 1P, PMe3 trans to Me). mer-(PMe3)3Ir(Me)(H)(BPin) (27). 1H NMR (CeDe) 5 -11.98 (dt, J = 131.9 Hz, 23.0 Hz, 1H, hydride), -0.06 (m, 3H, Me), 1.19 (s, 12H, 302C6H12), 1.14 (d, J = 29.9 Hz, 9H, PMe3 trans to hydride), 1.54 (t, J = 3.4 Hz, 18H, 2 PMe3 trans to each other). ”B NMR (C6D6) 6 38.6. 3"1>{‘H} NMR (C6D6) 6 - 57.8 (t, J= 22.9 Hz, 1P, PMe3 trans to hydride), -48.2 (d, J = 22.9 Hz, 2P, 2 PMe3 trans to each other). Anal. Calcd for C1611431rBO2P3: C, 34.11; H, 7.69. Found: C, 33.64; H, 7.70. mer-(PMe3)3Ir(BPin)(H)(Ph) (28). A solution of HBPin (55 mg, 0.43 mmol) in 2 mL pentane was added to a solution of (PMe3)3Ir(Ph) (194 mg, 0.39 mmol) in 5 mL 131 pentane. The reaction mixture was stirred at ambient temperature for 30 min and the solvent was removed under reduced pressure to give mer-(PMe3)31r(BPin)(H)(Ph) (241 mg, 95%). The product was recrystallized from a concentrated pentane solution at —30 °C to give colorless crystals. mp 118 °C (dec). IH NMR (C6D6, 25 °C) 8 -11.32 (dt, J = 131 Hz, 20 Hz, 1H, hydride), 1.16 (s, 12H, BO2C6H12), 1.41 (m, 27H, PMe3), 7.17-7.20 (m, 3H), 7.98 (br, 2H). 13C{1H} NMR (C6D6, 25 °C) 5 21.5 (d, J= 24.7 Hz), 22.2 (dt, J= 4.5 Hz, 19.1 Hz), 26.0 (s), 80.0 (s), 120.7 (s), 126.9 (s), 148.2 (br), 150.6 (br). 11B NMR (C613,) 6 35.8. 3‘P{'H} NMR (c6136) 6 —57.8 (t, J= 22.9 Hz, 1P), -45.6 (d, J: 22.0 Hz, 2P). Anal. Calcd for C21H45IrBO2P3: C, 40.32; H, 7.25. Found: C, 39.95; H, 7.38. fac—(PMe3)3Ir(BPin)(H)(SiEt3) (29). A solution of HSiEt3 (17 mg, 0.15 mmol) in 2 mL C6H6 was added to a solution of 18 (92 mg, 0.15 mmol) in 3 mL C6H6. The reaction mixture was stirred at ambient temperature for 2.5 days and the solvent was removed under reduced pressure to give colorless solid. The product was recrystallized from a concentrated pentane solution at —30 °C to give colorless crystals (83 mg, 86%). mp 134- 136 °C. ‘H NMR (C6D6) 6 -12.30 (dt, J: 117.0 Hz, 17.0 Hz, 1H, hydride), 0.84—0.95 (m, 3H, diastereotopic CH of CH2 groups of SiEt3), 1.25 (d, J = 6.7 Hz, 9H, PMe3 trans to BPin), 1.29 (s, 12H, BO2C6H12), 1.37 (d, J = 7.3 Hz, 9H, PMe3 trans to SiEt3), 1.46 (d, J = 7.6 Hz, 9H, PMe3 trans to hydride), 1.35-1.45 (m, 12H, diastereotopic CH of CH2 groups and CH3 groups of SiEt3). ”B NMR NMR (C613,) 6 36.3. ”C(‘m NMR (C60,) 5 10.7 (d, J= 1.9 Hz), 13.1 (dd, J= 5.8 Hz, 7.7 Hz), 24.6 (dt, J= 25.9 Hz, 4.8 Hz), 25.3 (dt, J = 22.1 Hz, 3.8 Hz), 25.4 (dt, J = 24.0 Hz, 3.8 Hz), 27.2 (s), 27.4 (5), 81.32 (s), 81.34 (s). 3'P{‘H} NMR (C6D6) 6 -66.4 (br, 1P, PMe3 trans to BPin), -64.2 (dd, J = 31.3 Hz, 132 19.8 Hz, 1P, PMe3 trans to SiEt3), -58.1 (dd, J = 19.8 Hz, 19.8 Hz, 1P, PMe; trans to hydride). Anal. Calcd for C21H551rBozP3Si: C, 38.00; H, 8.35. Found: C, 38.38; H, 8.52. mer-(PMe3)3Ir[B(NH)2C6H4](H)(Cl) (31). A solution of PMe; (102 mg, 1.34 mmol) in 2 mL THF was added dropwise to a solution of [Ir(COE)2Cl]2 (200 mg, 0.22 mmol) in 4 mL THF. The reaction mixture was stirred at room temperature for 2 h. The solvent was then removed under reduced pressure. The residue was redissolved in 6mL THF and H[B(NH)2C6H4] (51.7 mg, 0.44 mmol) was added to the mixture. The reaction mixture was stirred at ambient temperature for 12 h and then filtered through celite to remove trace suspension The filtrate was pumped down to give a white solid (226 mg, 88%). mp 170 °C (dec). ‘H NMR (CD202) 6 9.90 (dt, J = 138.2 Hz, 20.9 Hz, 1H, hydride), 1.53 (t, J = 3.6 Hz, 18H, 2 PMe3 trans to each other), 1.63 (dd, J = 7.5 Hz, 0.9 Hz, 9H, PMe3 trans to hydride), 5.83 (br, 2H, 2 NH), 6.64-6.66, 6.81-6.83 (m, 4H, C6H4). ‘3C{‘H} NMR(CD2C12) 6 19.3 (d, J= 26.9 Hz), 20.4 (dt, J: 4.3 Hz, 18.9 Hz), 108.7 (s), 117.5 (s), 139.0 (s). ”B NMR (CD202) 6 24.9. 3'P{‘H} NMR (CD202) 6 —48.2 (t, J = 20.8 Hz, 1P, PMe3 trans to hydride), —39.7 (d, J = 20.8 Hz, 2P, 2 PMe3 trans to each other). Anal. Calcd for C1 5H34IrBClN2P3: C, 31.40; H, 5.97; N, 4.88. Found: C, 31.34; H, 5.79; N, 4.89. mer-(PMe3)3Ir(BDAN)(H)(Cl) (32) and mer-(PMe3)3Ir(H)(BDAN)(Cl) (33). A solution of PMe3 (102 mg, 1.3 mmol) in 2 mL THF was added dropwise to a solution of [Ir(COE)2Cl]2 (200 mg, 0.22 mmol) in 4 mL THF. The reaction mixture was stirred at room temperature for 2 h. The solvent was then removed under reduced pressure. The residue was redissolved in 6 mL THF and HBDAN (74.5 mg, 0.44 mmol) was added to the mixture. The reaction mixture was stirred at ambient temperature for 12 h and was 133 then filtered through celite to remove trace suspension. The filtrate was pumped down to give an isomer mixture of 32 and 33 in a ratio of 90.4:9.6 (200 mg, 72%). mer- (PMe3)3Ir(BDAN)(H)(Cl) (32). 1H NMR (CDzClz) 8 -10.60 (dt, J = 137.5 Hz, 20.8 Hz, 1H, hydride), 1.63 (d, J = 6.4 Hz, 9H, PMe3 trans to hydride), 1.65 (t, J = 3.5 Hz, 18H, 2 PMC3 trans to each other), 5.49 (br, 2H, 2 NH), 6.19 (d, J = 7.5 Hz, 2H, BDAN), 6.86 (d, J: 8.5 Hz, 2H, BDAN), 7.03 (dd, J: 7.6 Hz, 8.0 Hz, 2H, BDAN). “B NMR(CD2C12) 6 28.0. ‘3C{‘H} NMR(CD2C12) 619.0 (d, J = 26.8 Hz), 20.3 (dt, J = 3.4 Hz, 18.9 Hz), 104.0 (s), 115.8 (s), 118.4 (s), 128.0 (s), 136.8 (s), 142.7 (d, J= 1.4 Hz). 3‘1){‘H} NMR (CDzClz) 5 —49.2 (t, J = 20.8 Hz, 1P, PMe3 trans to hydride), —39.7 (d, J = 20.8 Hz, 2P, 2 PMe3 trans to each other). mer-(PMe3)3Ir(H)(BDAN)(Cl) (33). 1H NMR (CDzClz) 8 - 23.97 (td, J = 16.0 Hz, 10.1 Hz, 1H, hydride), 1.50 (d, J = 6.6 Hz, 9H, PMe3 trans to BDAN), 1.56 (t, J = 3.3 Hz, 18H, 2 PMe3 trans to each other), 5.74 (br, 2H, 2 NH), 6.18 (d, J= 7.3 Hz, 2H, BDAN), 6.82 (dd, J= 8.7 Hz, 12.4 Hz, 2H, BDAN), 6.99 (d, J= 8.0 Hz, 2H, BDAN). “B NMR (CD2C12) 6 38.6. 3‘P{1H} NMR(CD2C12) 6 —52.4 (br, 1P, PMe3 trans to BDAN), —43.1 (d, J = 26.9 Hz, 2P, 2 PMe3 trans to each other). Anal. Calcd for C19H361rBC1N2P3: C, 36.58; H, 5.82; N, 4.49. Found: C, 36.62; H, 5.87; N, 4.43. mer-(PMe3)3Ir[B(NMe)2C6H4](H)(Cl) (34) and mer-(PMe3)3Ir(H)[BG‘JMe)2C6 H4](Cl) (35). A solution of PMe3 (102 mg, 1.3 mmol) in 2 mL THF was added dropwise to a solution of [Ir(COE)2Cl]2 (200 mg, 0.22 mmol) in 4 mL THF. The reaction mixture was stirred at room temperature for 2 h. The solvent was then removed under reduced pressure. The residue was redissolved in 6 mL THF and H[B(NMe)2C6H4] (53.7 mg, 0.37 mmol) was added to the mixture. The reaction mixture was stirred at ambient temperature 134 for 12 h and then filtered through celite to remove trace suspension. The filtrate was pumped down to give an isomer mixture of 34 and 35 in a ratio of 14.42856 (220 mg, 82%). mer-(PMe3)3Ir[B(NMe)2C6H4](H)(Cl) (34). ‘H NMR(CD2C12) 6 -1075 (dt, J = 139.0 Hz, 19.6 Hz, 1H, hydride), 1.47 (t, J = 3.5 Hz, 18H, 2 PMe3 trans to each other), 1.68 (dd, J = 6.9 Hz, 0.8 Hz, 9H, PMe3 trans to hydride), 3.28 (s, 3H, Me), 3.55 (s, 3H, Me), 6.77-6.85 (m, 4H, C6114). “B NMR (CD2C12) 6 28.6. 3‘P{‘H} NMR (CDZClz) 5 —54.9 (t, J = 22.0 Hz, 1P, PMe3 trans to hydride), —39.4 (d, J = 22.0 Hz, 2P, 2 PMe3 trans to each other). mer-(PMe3)3Ir(H)[B(1\lMe)2C6H4](Cl) (35). 1H NMR (CDzClz) 8 - 23.97 (td, J = 14.3 Hz, 12.2 Hz, 1H, hydride), 1.38 (t, J= 3.3 Hz, 18H, 2 PMe3 trans to each other), 1.54 (d, J = 6.7 Hz, 9H, PMe3 trans to [B(NMe)2C6H4], 3.35 (s, 3H, Me), 3.68 (s, 3H, Me), 6.77-6.85 (m, 4H, C6H4). ”B NMR(CD2C12) 6 38.3. '3C(‘H} NMR (CDzClz) 5 20.0 (dt, J = 5.5 Hz, 18.9 Hz), 20.3 (d, J = 22.0 Hz), 32.1 (s), 32.8 (s), 106.2 (s), 106.6 (s), 117.0 (s), 117.2 (s), 140.6 (d, J = 3.4 Hz), 141.8 (d, J = 4.8 Hz). 3‘P(‘H} NMR(CD2C12) 5 —50.6 (br, 1P, PMe; trans to [B(NMe)2C6H4], -40.6 (d, J = 26.9 Hz, 2P, 2 PMe3 trans to each other). Anal. Calcd for C17H33IrBClN2P3: C, 33.92; H, 6.36; N, 4.66. Found: C, 34.10; H, 6.35; N, 4.57. Screening Experiments The general procedure for the synthesis of phenylboronate esters, catalyzed by 13 or 14/a monodentate phosphine ligand, is illustrated by the following example. Decane (0.626 M in benzene, 50 11L, 0.031 mmol), HBPin (51 11L, 0.351 mmol) were charged into a J. Young NMR tube. 13 (3 mg, 7.3 x 10'3 mmol) was charged into a GC-vial and 135 dissolved in C6H6 (150 uL). PMe3 (1.5 uL, 0.015 mmol) was added into the solution of 13 via a microsyringe. The mixture was then transferred into the J. Young NMR tube. Benzene (150 uL x 2) was used to wash the residue to the J. Young NMR tube. The reaction mixture was heated at 150 °C and monitored by HB NMR spectra. After HBPin was consumed, an aliquot of the reaction mixture was diluted with CHzClz and a GC-FID chromatogram was obtained. From the calibration curve of PhBPin vs. decane, GC yield of PhBPin formation was obtained. The results of borylation of benzene catalyzed by 13 or 14/a monodentate phosphine ligand are summarized in Table 3. The general procedure for the synthesis of phenylboronate esters, catalyzed by 13/a chelating phosphine ligand, is illustrated by the following example. Decane (0.626 M in benzene, 50 111., 0.031 mmol), HBPin (51 11L, 0.351 mmol) were charged into a J. Young NMR tube. 13 (2.9 mg, 7.0 x 10‘3 mmol) and dppe (2.8 mg, 7.0 x 10'3 mmol) were charged into two separate GC-vials and dissolved in C6H6 (150 uL x 2). The mixture was then transferred into the J. Young NMR tube. Benzene (150 11L) was used to wash the residue to the J. Young NMR tube. The reaction mixture was heated at 100 0C or 150 °C and monitored by 11B NMR spectra. Afier HBPin was consumed, an aliquot of the reaction mixture was diluted with CHzClz and a GC-FID chromatogram was obtained. From the calibration curve of PhBPin vs. decane, GC yield of PhBPin formation was obtained. The results of borylation of benzene catalyzed by l3/a chelating phosphine ligand are summarized in Table 4. The general procedure for the synthesis of phenylboronate esters, catalyzed by 13/a nitrogen or oxygen or sulfur containing ligand, is illustrated by the following example. Dodecane (0.471 M in benzene, 50 1.1L, 0.024 mmol), HBPin (51 uL, 0.351 136 mmol) were charged into a J. Young NMR tube. 13 (2.9 mg, 7.0 x 10'3 mmol) and bpy (1.1 mg, 7.0 x 10’3 mmol) were charged into two separate GC-vials and dissolved in C6H6 (150 uL x 2). The mixture was then transferred into the J. Young NMR tube. Benzene (150 uL) was used to wash the residue to the J. Young NMR tube. The reaction mixture was heated at 100 °C or 150 0C and monitored by HB NMR spectra. After HBPin was consumed, an aliquot of the reaction mixture was diluted with CH2C12 and a GC-FID chromatogram was obtained. From the calibration curve of PhBPin vs. dodecane, GC yield of PhBPin formation was obtained. The results of borylation of benzene catalyzed by 13/a nitrogen or oxygen or sulfur containing ligand are summarized in Table 5. NMR Tube Reactions Metathesis reaction between Cp*Ir(PMe3)(Ph)(H) (5) and HBPin in Cans. Cp*Ir(PMe3)(Ph)(H) (9 mg, 0.019 mmol) and HBPin (28.7 mg, 0.224 mmol) were dissolved in C6D6 (540 uL) and the mixture was transferred to a J. Young NMR tube. The reaction mixture was heated at 150 °C in an oil bath and monitored by 1H, HB, and 3'P{‘H} NMR. Thermolysis of Cp*Ir(PMe3)(I-I)(BPin) (l) in C6D6. Compound 1 (10 mg, 0.019 mmol) dissolved in C6D6 (550 11L) was transferred to a J. Young NMR tube. The solution was heated at 200 °C in an oil bath for 2 weeks and 280 °C for 1 day. The reaction was monitored by 1H, llB, and 3|P{1H} NMR. Metathesis reaction between Cp*Rh(PMe3)(Ph)(H) (4) and HBPin in C6D6. Cp*Rh(PMe3)(Ph)(H) (10 mg, 0.026 mmol) and HBPin (39 mg, 0.305 mmol) were 137 dissolved in C6D6 (550 11L) and the mixture was transferred to a J. Young NMR tube. The reaction mixture was heated at 95 °C in an oil bath and monitored by IH, HB, and 3'P('H} NMR. Crossover experiment. Cp*Ir(P(CD3)3)(H)2 (8) (10 mg, 0.024 mmol), (C5Me4Et)Ir(PMe3)(H)2 (9) (10.1 mg, 0.024 mmol), and HBPin (15.5 mg, 0.121 mmol) were dissolved in C6H6 (550 11L) and the reaction mixture was transferred to a J. Young NMR tube. The reaction mixture was heated at 150 °C in an oil bath and monitored by HB NMR. Afier HBPin was consumed, an aliquot of the reaction mixture was diluted with CHzClz and a GC-MS chromatogram was obtained. From the chromatogram of the crude mixture there was no crossover products observed. Therefore, crossover during catalytic borylation was minimal. Anisole borylation catalyzed by Cp*IrH4 (11). Cp*IrH4 (11) (10 mg, 0.030 mmol) and HBPin (23 mg, 0.180 mmol) were dissolved in a 1:2 (V/V) anisole/cyclohexane solution (500 uL). The reaction mixture was transferred to a J. Young NMR tube, heated at 150 °C in an oil bath, and monitored by I'B NMR. After HBPin was consumed, an aliquot of the reaction mixture was diluted with CHzClz and a GOP ID chromatogram was obtained. The isomer ratio of C6H4(0Me)(BPin) (o:m:p) was determined to be 3:49:48. Thermolysis of (MesH)lr(BPin)3 (14) in C6116. (MesH)Ir(BPin)3 (14) (15.2 mg, 0.022 mmol) dissolved in C6H6 (550 11L) was transferred to a J. Young NMR tube. The solution was heated at 150 °C in an oil bath and monitored by HB NMR. After complex 14 was consumed, an internal standard solution (decane/CoHo) was added to the reaction mixture. An aliquot of the mixture was diluted with CHzClz and a GC-F ID chromatogram 138 was obtained. Thermolysis of 14 in C6H6 produced 3 equivalents of PhBPin and black iridium metal. 14 + excess HBPin in Cng. Compound 14 (10 mg, 0.014 mmol) and HBPin (11 mg, 0.086 mmol) dissolved in C6D6 (550 11L) was transferred to a J. Young NMR tube. The reaction mixture was heated at 150 °C in an oil bath and monitored by HB NMR. After 5 h at 150 °C, the reaction led to decomposition and PinB-O-BPin. No C6D5BP1n was observed. 14 + excess HBPin in C5D; in the presence of 2 equiv. of PMe3. Compound 14 (10 mg, 0.014 mmol) and HBPin (11 mg, 0.086 mmol) dissolved in C6D6 (550 11L) was transferred to a J. Young NMR tube. PMe; (3 11L, 0.029 mmol) was added to the NMR tube via a microsyringe. The reaction mixture was heated at 150 °C in an oil bath and monitored by 1'B NMR. After 5 h at 150 °C, HBPin was completely converted to PhBPin. 14 + 2 equiv. PMe; in toluene-d3. Compound 14 (10 mg, 0.014 mmol) dissolved in toluene-d3 (500 1.1L) was transferred to a J. Young NMR tube. PMe3 (3 11L, 0.029 mmol) was added to the NMR tube via a microsyringe. The reaction was monitored by 1H, HB, and 31P{1H} NMR at room temperature. The reaction generated a mixture of compound 25 and (n6-C7Dg)1r(BPin)3 in a 2:1 ratio. 37 + HBPin in C6D5. (PMe3)4Ir(H) (37) (15 mg, 0.030 mmol) dissolved in C6D6 (332 11L) in a GC vial was transferred to a J. Young NMR tube. Additional C6D6 (166 11L) was used to wash the residue into the NMR tube. HBPin (4.4 11L, 0.030 mmol) was added into the NMR tube via a microsyringe. At room temperature, the starting material was gradually converted into a mixture of mer,cis-(PMe3)3Ir(H)2(BPin) (38) and fac- 139 (PMe3)3Ir(H)2(BPin) (36). The sample was allowed to stand at room temperature for 6 days to give compound 36 as the predominant species. mer,cis-(PMe3)3Ir(H)2(BPin) (38). lH NMR (CsDo) 6 -12.18 (dt, J: 114.7 Hz, 23.2 Hz, 1H, hydride trans to PMeg), -10.46 (q, 1H, hydride trans to BPin), 1.21 (s, 12H, B02C6H12), 1.49 (d, 9H, PMe; trans to hydride), 1.69 (t, J = 3.5 Hz, 18H, 2 PMe; trans to each other). 1‘B NMR (Cng) 5 38.6. 3‘1>{‘H} NMR (Cng) 6 -58.1 (t, J = 22.6 Hz, 1?, PMe3 trans to hydride), -48.1 (d, J = 22.6 Hz, 2P, 2 PMe3 trans to each other). fac-(PMe3)3Ir(H)2(BPin) (36). 1H NMR (Cng) 5 -11.83 (symmetrical second order m, 2H, hydride), 1.25 (s, 12H, BOngHrz), 1.32 (d, J = 7.0 Hz, 9H, PMC3 trans to BPin), 1.69 (d, J = 7.6 Hz, 18H, 2 PMe3 trans to hydride). ”B NMR (C6D6) 6 38.6. 3‘1>(‘H} NMR (C6D6) 6 -62.0 (br, 1P, PMe3 trans to BPin), - 54.59 (d, J = 23.2 Hz, 2P, 2 PMe3 trans to hydride). 37 + BzPinz in C6D6. B2Pin2 (7.9 mg, 0.031 mmol) dissolved in C6D6 (166 11L) was transferred to a J. Young NMR tube which was charged with (PMe3)4Ir(H) (37) (15.4 mg, 0.031 mmol) in C6D6 (166 11L). Additional CoDo (166 1.1L x 2) was used to wash the residue into the NMR tube. The reaction mixture was heated at 60 °C and monitored by lH, HB, and 3'P{1H} NMR spectra. The starting material was gradually converted into a mixture of mer,trans-(PMe3)3Ir(BPin)2(H) (20) and fac-(PMe3)3Ir(BPin)2(H) (21). fac- (PMe3)3Ir(BPin)2(H) (21) was the major species after the temperature was increased to 100 °C for 7 h. mer,trans-(PMe3)3Ir(BPin)2(H) (20). 1H NMR (C6D6) 5 -12.36 (dt, J = 117.0 Hz, 21.7 Hz, 1H, hydride trans to PMe3), 1.22 (s, 12H, B02C6H12), 1.49 (d, J = 8.0 Hz, 9H, PMe3 trans to hydride), 1.74 (t, J = 3.4 Hz, 18H, 2 PMe3 trans to each other). “B NMR (e.u.) 6 38.9. 3'P{‘H} NMR (Cst) 6 -59.6 (t, J = 22.0 Hz, 1P, PMe3 trans to hydride), -50.8 (d, J = 22.0 Hz, 2P, 2 PMe3 trans to each other). fac-(PMe3)3Ir(BPin)2(H) 140 (21). mp 120-122 °C. 1H NMR (CoDo) 5 -11.66 (dt, J= 118.1 Hz, 18.1 Hz, 1H, hydride trans to PMe3), 1.29 (s, 24H, B02C6H12), 1.41 (W, 18H, 2 PMe3 trans to BPin), 1.58 (d, J = 8.0 Hz, 9H, PMe3 trans to hydride). '3C{‘H} NMR (C6D6) 6 23.7 (dt, J: 26.9 Hz, 5.3 Hz), 25.1 (br), 25.8 (s), 26.3 (s), 80.3 (s). “B NMR (C6D6) 6 38.6. 3‘1>{‘H} NMR (C6D6) 5 -61.8 (br, 2P, 2 PMe3 trans to BPin), -56.6 (t, J = 22.0 Hz, 1P, PMe3 trans to hydride). fac-(PMe3)3Ir(BPin)2(H) (21) was independently synthesized and isolated in 80 % yield. Anal. Calcd for C21H521rB204P3: C, 37.34; H, 7.76. Found: C, 37.35; H, 7.72. 18 + dppe in C6D6. Dppe (8 mg, 0.020 mmol) dissolved in C6D6 (166 uL) was transferred to a J. Young NMR tube, which was charged with 18 (12.5 mg, 0.020 mmol) in C6D6 ( 166 11L). Additional CoDo (166 11L) was used to wash the residue into the NMR tube. The reaction mixture was allowed to stand at room temperature for 3 days to give Ir(PMe3)2(dppe)(BPin) (19) as the predominant species. 1H NMR (C6D6) 6 1.10 (s, 12H, BOzCsHiz), 1.33 (t, J= 3.3 Hz, 18H, 2 PMe3), 1.92-2.18 (m, 4H, CH2), 6.98-7.12, 7.16- 7.28, 7.72-7.89, 7.91-7.98 (m, 20H, phenyl groups of dppe). “B NMR (C,D,) 6 38.8. 3‘1>(‘H} NMR (Cst) 6 —58.9 (dd, J = 141.6 Hz, 26.8 Hz, 2P, 2 PMe3), 39.1 (dt, J = 141.6 Hz, 13.4 Hz, 1P, PPh2 cis to BPin), 46.1 (br, 1P, Pth trans to BPin). 18 + HBPin in C6D6. Compound 18 (12.8 mg, 0.021 mmol) dissolved in C6D6 (500 pL) was transferred to a J. Young NMR tube. HBPin (3 uL, 0.021 mmol) was added to the NMR tube via a microsyringe. The reaction was monitored by 1H, 1'B, and 31P{1H} NMR at room temperature. mer,trans-(PMe3)3Ir(BPin)2(H) (20) was the initial predominant species, and it gradually isomerized to fac-(PMe3)3Ir(BPin)2(H) (21) after heating at 70 °C for 11h. 141 18 + ClBCat in C6D6. Compound 18 (12.1 mg, 0.019 mmol) and ClBCat (3 mg, 0.019 mmol) dissolved in C6D6 (500 uL) were transferred to a J. Young NMR tube. The reaction was monitored by 1H, 11B, and 31P{1H} NMR at room temperature. After 3 days at room temperature, the predominant species in the reaction mixture was mer- (PMe3)3Ir(BPin)(BCat)(Cl) (22) with BPin group trans to Cl. 1H NMR (c613,) 6 1.18 (s, 12H, BOZCOH12). 1.29 (d, J = 7.3 Hz, 9H, PMC3 trans to BCat group), 1.47 (t, J = 3.7 Hz, 18H, 2 PMe3 trans to each other), 6.84, 7.19 (AA'BB', 4H, BCat). "B NMR (C613,) 6 28.1, 41.6. 3‘1>{'H} NMR (C613,) 6 —56.2 (br, PMe3 trans to BCat), —39.3 (d, J: 29.3 Hz, 2 PMe3 trans to each other). Thermolysis of 18 in C6D6. Compound 18 (10.2 mg, 0.016 mmol) dissolved in C6D6 (500 11L) was transferred to a J. Young NMR tube. The reaction mixture was heated at 100 °c in an oil bath and monitored by 1H, ”B, and 3‘1>{‘H} NMR. After 38 h at 100 °C, complex 18 was converted to (PMe3)4Ir(D) and C6D5BPin. Thermolysis of 25 in C6D6. Compound 25 (15 mg, 0.019 mmol) dissolved in C6D6 (500 11L) was transferred to a J. Young NMR tube. The reaction mixture was heated at 150 °C in an oil bath and monitored by 1H, 1'3, and 3'P{1H} NMR. The reaction gave fac-(PMe3)3Ir(D)3 as the final iridium containing product and generated 3 equiv. of CgDsBPin. Thermolysis of 18 in C5H51. Complex 18 (12 mg, 0.019 mmol) dissolved in C6H51 (500 uL) was transferred to a J. Young NMR tube. The reaction resulted in immediate white precipitation. The reaction mixture was heated at 100 °C for 1 h in an oil bath. No isomer mixture of C6H4(I)(BPin) was detected by GC-FID. 142 Thermolysis of 25 in C6H51. Compound 25 (11.6 mg, 0.014 mmol) dissolved in C6H51 (500 uL) was transferred to a J. Young NMR tube. The reaction mixture was heated at 150 °C for 29 h in an oil bath. The reaction produced m- and p-C6P14(I)(BPin) in 54% GC yield, in addition to a 45% yield of PhBPin. Kinetic Isotope Effect Experiments Catalytic borylation in a molar ratio 1:1 mixture of C6H6/C5D6 with the Irl pro-catalyst (2 mol% 13 and 4 mol% PMe3). A solution of (Ind)Ir(COD) (13) (6 mg, 0.014 mmol) in C6H6/C6D6 (1:1) (175 uL x 2) was mixed with a solution of PMe3 (3 11L) in Cng/Cng (1 :1) (175 uL x 2). The solution mixture was then transferred to a J. Young NMR tube. Additional CgHb/Cng (1:1) (175 11L x 4) was used to wash the residue into the NMR tube. HBPin (104 11L, 0.717 mmol) was added to the NMR tube via an auto- pipette. The reaction mixture was heated at 150 °C in a constant temperature oil bath (Cole-Parmer Polystat Constant Temperature Circulator). The reaction was monitored by HB NMR. The ratio of C6D5BPin : C6H5BPin determined from GC-FID after calibration was 1.00:2.29. Catalytic borylation in a molar ratio 1:1 mixture of C6H6/C5D6 with the Ir"l pre—catalyst (2 mol% 14 and 4 mol% PMe3). A solution of (MesH)Ir(BPin)3 (14) (10 mg, 0.014 mmol) in C6H6/C6D6 (1:1) (175 11L x 2) was mixed with a solution of PMe3 (3 uL) in CgHg/Cng (1:1) (175 11L x 2). The solution mixture was then transferred to a J. Young NMR tube. Additional CgHg/Cng (1:1) (175 uL x 4) was used to wash the residue into the NMR tube. HBPin (104 1.1L, 0.717 mmol) was added to the NMR tube via 143 an auto-pipette. The reaction mixture was heated at 150 °C in a constant temperature oil bath (Cole-Farmer Polystat Constant Temperature Circulator). The reaction was monitored by 11B NMR. The ratio of C6D5BPin : C6H5BPin determined from GC-F ID after calibration was 1.00:2.28. Catalytic borylation in 1,3,5-C6D3H3 with the Irl pre-catalyst (2 mol% l3 and 4 mol% PMe3). A solution of 13 (3 mg, 0.007 mmol) in 1,3,5-C6D3H3 (166 11L) was mixed with a solution of PMe3 (1.5 uL) in 1,3,5-C6D3H3 (166 11L). The solution mixture was then transferred to a J. Young NMR tube. Additional 1,3,5-C6D3H3 (166 uL) was used to wash the residue into the NMR tube. HBPin (52 11L, 0.358 mmol) was added to the NMR tube via an auto-pipette. The reaction mixture was heated at 150 0C in a constant temperature oil bath. The reaction was monitored by HB NMR. The ratio of 1,3,5-C6D2H3(BPin) : 1,3,5-C6D3H2(BPin) determined by the ‘H NMR of the crude mixture was 1.00:2.06. Catalytic borylation in 1,3,5-C6D3H3 with the Ir"1 pre-catalyst (2 mol% 14 and 4 mol% PMC3). A solution of 14 (5 mg, 0.007 mmol) in 1,3,5-C6D3H3 (166 11L) was mixed with a solution of PMe3 (1.5 11L) in 1,3,5-C6D3H3 (166 11L). The solution mixture was then transferred to a J. Young NMR tube. Additional 1,3,5-C6D3H3 (166 uL) was used to wash the residue into the NMR tube. HBPin (52 uL, 0.358 mmol) was added to the NMR tube via an auto-pipette. The reaction mixture was heated at 150 °C in a constant temperature oil bath. The reaction was monitored by 1'B NMR. The ratio of 1,3,5-C6D2H3(BPin) : 1,3,5-C6D3H2(BPin) determined from the 1H NMR of the crude mixture was 1.00:1.94. 144 Thermolysis of 18 in a molar ratio 1:1 mixture of C6H6/C6D6. Compound 18 (25 mg, 0.060 mmol) dissolved in CgHg/CoDo (1:1) (166 11L x 2) was transferred to a J. Young NMR tube. Additional CgHg/CgDr, (1:1) (166 11L) was used to wash the residue into the NMR tube. The reaction mixture was heated at 150 °C in an oil bath and monitored by 11B and 3 IP{'H} NMR. The ratio of C6D5BPin : C6H5BPin determined from GC-F ID after calibration was 1.00:2.67. Thermolysis of 25 in a molar ratio 1:1 mixture of C6H6/C6D6. Compound 25 (32 mg, 0.040 mmol) dissolved in C6H6/C6D6 (1:1) (166 11L x 2) was transferred to a J. Young NMR tube. Additional C6H6/C6D6 (1:1) (166 11L) was used to wash the residue into the NMR tube. The reaction mixture was heated at 150 °C in an oil bath and monitored by 11B and 3 lP{1H} NMR. The ratio of C6D5BPin : C6H5BPin determined from GC-FID after calibration was 1.00:2.53. Thermolysis of 18 in 1,3,5-C6D3H3. Compound 18 (25 mg, 0.060 mmol) dissolved in 1,3,5-C6D3H3 (166 11L x 2) was transferred to a J. Young NMR tube. Additional 1,3,5-C6D3H3 (166 uL) was used to wash the residue into the NMR tube. The reaction mixture was heated at 150 °C in an oil bath and monitored by 11B and 3 1P{‘H} NMR. The ratio of 1,3,5-C6D2H3(BPin) : 1,3,5-C6D3H2(BPin) determined from the lH NMR of the crude mixture was 1.00:2.37. Thermolysis of 25 in 1,3,5-C6D3H3. Compound 25 (32 mg, 0.040 mmol) dissolved in 1,3,5-C6D3H3 (166 uL x 2) and transferred to a J. Young NMR tube. Additional 1,3,5-C6D3H3 (166 11L) was used to wash the residue into the NMR tube. The reaction mixture was heated at 150 °C in an oil bath and monitored by 1'B and 31P{'H} 145 NMR. The ratio of 1,3,5-C6D2H3(BPin) : 1,3,5-C6D3H2(BPin) determined from the 1H NMR of the crude mixture was 1.00:1.93. Arylboronate Ester Syntheses The general procedure for the synthesis of arylboronate esters, catalyzed by solutions of compound 1, is illustrated for the synthesis of 1,3,5-C6H3(CF3)2(BPin). Compound 1, (80 mg, 0.15 mmol) and HBPin (96 mg, 0.75 mmol) were dissolved in 4 mL 1,3-bis(trifluoromethyl)benzene and heated at 150 °C in a constant temperature circulator for 10 hours in a thick—walled, air—free flask. The solution was then transferred to a vial and the solvent was removed under vacuum at room temperature. The residue was chromatographed on a silica gel column, eluting with CHzClz, to yield 1,3,5- C6H3(CF3)2(BPin) as a colorless solid (182 mg, 81% based on HBPin). mp (65-66 °C). ‘H NMR (CDC13) 6 1.35 (s, 12H, 302C6H12), 7.92 (s, 1H), 8.22 (s, 2H). ‘3C{‘H} NMR (CDC13) 5 24.8, 84.9, 123.5 (J = 272.4 Hz, 2C), 124.7, 130.9 (J = 32.7 Hz, 2C), 134.7 (2C). ”3 NMR(CDC13) 6 30. ‘91: NMR(CDC13)5 —63. Anal. Calcd for C14H15BF602: C, 49.44; H, 4.45. Found: C, 49.55; H, 4.53. GC-MS (m/z) 340. The general procedure for the synthesis of arylboronate esters, catalyzed by solutions of compound 1, generated in situ from compound 2, is illustrated for the synthesis of CgHaMe(BPin). Compound 2, (70 mg, 0.17 mmol) and HBPin (133 mg, 1.04 mmol) were dissolved in 4 mL toluene and heated at 150 °C in a constant temperature circulator for 10 hours in a thick—walled, air—free flask. The solution was then transferred to a vial and the solvent was removed under vacuum at room temperature. The residue 146 was chromatographed on a silica gel column, eluting with CHzClz, to yield C6HaMe(BPin). The reaction gave 3 isomers, m-CgHaMe(BPin) : p-CgHaMe(BPin) : 0- C6H4Me(BPin) in the ratio of 62:34:4 (173 mg, 91% based on HBPin). The identity of o- C6HaMe(BPin) and p-C6H4Me(BPin) were established by comparing spectroscopic data to those in the literature,25 and by comparing GC-MS data for the mixture to data for the independently prepared pure isomers. The identity of m-C6H4Me(BPin) was confirmed by independent synthesis of an authentic sample using the literature method.2 m- CgHaMe(BPin) was isolated as a colorless solid. mp 34-35 °C. 1H NMR (CDC13) 5 1.33 (s, 12H, BOzC6H12), 2.34 (s, 3H, Me), 7.25 (m, 2H), 7.59 (m, 1H), 7.62 (s, 1H). ”B NMR (CDCl3) 5 30.7. Anal. Calcd for C13H19B02: C, 71.59; H, 8.78. Found: C, 71.36; H, 9.33. GC-MS (m/z) 218. The general procedure for the synthesis of arylboronate esters, catalyzed by solutions of compound 3, is illustrated for the synthesis of 1,3,5-C6H3(CF3)2(BPin). Compound 3, (5 mg, 0.013 mmol) and HBPin (90 mg, 0.70 mmol) were dissolved in 550 11L 1,3-bis(trifluoromethyl)benzene and heated at 150 °C in a constant temperature circulator for 3 hours in a J. Young NMR tube. The solution was transferred to a vial and the solvent removed under vacuum at room temperature. The residue was chromatographed on a silica gel column, eluting with CHzClz, to yield 1,3,5— C6H3(CF3)2(BPin) as a colorless solid (203 mg, 86% based on HBPin). F F F BPin F F C6F5(BPin). Catalytic addition of HBPin to C6HF 5 using solutions of compounds 1 (generated in situ from compound 2) and 3 gave C6F5(BPin) as a colorless solid (205 147 mg, 81% based on HBPin, and 85 mg, 41% based on HBPin, for 1 and 3, respectively). mp 35-36 °C. 1H NMR(CDC13) 6 1.36 (s, 12H, BOngHrz). ”B NMR (CDC13) 6 29. "’1: NMR (CDCl,) 6 -129.5 (m, 2F), —149.7 (m, 1F),—161.9(m, 2F). Anal. Calcd for CtzleBFstz C, 49.02; H, 4.11. Found: C, 48.33; H, 4.59. GC-MS (m/z) 294. C6H4(CF3)(BPin) (isomer mixture). Catalytic addition of HBPin to benzotrifluoride using solutions of 1 gave 2 isomers, m-C6Ha(CF3)(BPin) : p- C6H4(CF3)(BPin) in a 2:1 ratio (202 mg, 99% based on HBPin). HBPin addition catalyzed by solutions of 3 gave m-C6Ha(CF3)(BPin) : p-C6Ha(CF3)(BPin) in a 2:1 ratio (161 mg, 84% based on HBPin). The proton chemical shifts of the two isomers were determined by selective decoupling of peaks in the aromatic region. m-C6H4(CF3)(BPin). lH NMR (CDCl3) 5 1.34 (s, 12H, BOzCoth), 7.47 (t, 1H), 7.68 (d, 1H), 7.96 (d, 1H), 8.05 (s, 1H). “B NMR (CDC13) 6 30.2. ”F NMR(CDC13) 6 —62.9. p-C6Ha(CF3)(BPin). lH NMR (CDCl;) 6 1.34 (s, 12H, 302C6H12), 7.59 (d, 2H), 7.89 (d, 2H). “B NMR (CDCl3) 5 30.2. 19F NMR (CDC13) 5 -63.3. Anal. Calcd for C13H16BF302: C, 57.39; H, 5.93. Found: C, 57.48; H, 6.40. GC-MS (m/z) 272. C6H4Me(BPin) (isomer mixture). Catalytic addition of HBPin to toluene using solutions of 3 gave 3 isomers, m-C6HMe(BPin) : p-C6H4Me(BPin) : 0-C6H4Me(BPin) in a 63:32:5 ratio (110 mg, 72% based on HBPin). 148 Me C6H4(0Me)(BPin) (isomer mixture). Catalytic addition of HBPin to anisole using solutions of 3 gave 3 isomers, m-C6H4(0Me)(BPin) : p-C6H4(0Me)(BPin) : 0- C6Ha(OMe)(BPin) in a 67:25:8 ratio (106 mg, 65% based on HBPin). The proton chemical shifts of the 2 major isomers were determined by selective decoupling experiments and by comparison to spectra of independently prepared authentic samples. m-C6H4(0Me)(BPin). 'H NMR (CDC13) 6 1.33 (s, 12H, BOszth), 3.82 (s, 3H, OMe), 7.25-7.44 (m, 3H), 7.00 (m, 1H). ”B NMR(CDC13) 6 30.8. p-CoHrt(0Me)(BPin). 1H NMR (CDC13) 5 1.31 (s, 12H, BOzCéle), 3.81 (s, 3H, OMe), 6.88 (d, 2H), 7.74 (d, 2H). “B NMR (CDC13) 6 30.8. Anal. Calcd for c,,H,9Bo3: C, 66.70; H, 8.18. Found: C, 66.69; H, 8.50. GC-MS (m/z) 234. QNMez BPin C6H4(NMe2)(BPin) (isomer mixture). Catalytic addition of HBPin to N,N- dimethylaniline using solutions of 3 gave m-C6Ha(NMe2)(BPin) : p-C6Ha(NMe2)(BPin) : 0-C6H4(NMe2)(BPin) in the ratio 55:43:2 (113 mg, 65% based on HBPin). The proton chemical shifts of p-C6Ha(NMe2)(BPin) and m—C6Ha(NMe2)(BPin) isomers were determined by selective decoupling and NOE experiments. Resonances in the aromatic region were assigned to m- and p-CgHa(NMe2)(BPin) isomers by selective decoupling experiments. Resonances for the methyl groups of We; in m- and p—C6Ha(NMe2)(BPin) were assigned to each isomer based on one-dimensional NOE experiments. In the NOE experiment, the methyl resonances of the NMez groups were irradiated in order to establish through space relationships with their respective aromatic protons. Irradiation of 149 the methyl resonance at 5 2.97 resulted in an enhancement of the peak at 56.68, corresponding to the aromatic protons assigned to p-C6Ha(NMe2)(BPin). Irradiation of the second methyl resonance at 5 2.95 resulted in enhancement of the peaks at 5 6.85 and 5 7.21, corresponding to aromatic protons assigned to m-C6H4(NMe2)(BPin). The integration of the methyl resonances of We; by deconvolution of the peaks, indicates a higher percentage of the meta isomer with respect to the para isomer. From the integration of the 1H NMR spectrum, the peaks in the GC-MS were assigned accordingly to each isomer. m-C6H4(NMe2)(BPin). 1H NMR (CDC13) 5 1.33 (s, 12H, B02C6H12), 2.95 (s, 6H, N (CH3)2), 6.85 (m, 1H), 7.17-7.28 (m, 3H). “13 NMR(CDC13) 5 31.1. p-C6H4(NM62)(BPin). 1H NMR(CDC13) 5 1.32 (s, 12H, BOzCole), 2.97 (s, 6H, NMez), 6.68 (d, 2H), 7.69 (d, 2H). ”3 NMR (CDCl;,) 6 31.1. Anal. Calcd for C14H22BN02: C, 68.04; H, 8.97. Found: C, 67.84; H, 9.11. GC-MS (m/z) 247. BPin C6H4(CHMe2)(BPin) (isomer mixture). Catalytic addition of HBPin to cumene using solutions of 3 gave 3 isomers, m-C6H4(CHMez)(BPin) : p-C6H4(CHMe2)(BPin) : o- C6H4(CHMez)(BPin) in a 66:33:] ratio (115 mg, 67% based on HBPin). The proton chemical shifts of the 2 major isomers were determined by selective decoupling experiments. m-C6H4(CHMe2)(BPin). lH NMR (CDC13) 5 1.24 (d, 6H, 2Me), 1.33 (s, 12H, BOzCsle). 2.91 (m, 1H, CH), 7.27-7.33 (m, 2H), 7.62 (d, 1H), 7.65 (s, 1H). HB NMR (CDC13) 6 31.1. p-CsHa(CHMe2)(BPin). 1H NMR (CDClg) 61.23 ((1, 6H, 2Me), 1.32 (s, 12H, BOzCole), 2.91 (m, 1H, CH), 7.22 (d, 2H), 7.73 (d, 2H). “B NMR (CDC13) 6 31.1. GC-MS (m/z) 246. 150 Me Me BPin 1,3,5-C6H3Me2(BPin) Catalytic addition of HBPin to m-xylene using solutions of 3 gave 1 major product. 1,3,5-C6H3Me2(BPin) was isolated as a white solid (119 mg, 73% based on HBPin). mp 90-91 0C. lH NMR (CDC13) 6 1.33 (s, 12H, BozC,H.2), 2.30 (s, 6H, 2Me), 7.09 (s, 1H), 7.42 (s, 2H). l3C(‘H} NMR (CDC13) 6 21.1, 24.8, 83.6, 132.4, 132.9, 137.1. ”B NMR (CDC13) 6 30.9. Anal. Calcd for CraHerOZ: C, 72.44; H, 9.12. Found: C, 72.38; H, 9.44. GC-MS (m/z) 232. Me N Me \ I / BPin 2,4,6-C5NH2Mez(BPin) Catalytic addition of HBPin to 2,6-lutidine using solutions of 3 gave 2,4,6-C5NH2Me2(BPin), isolated as colorless crystals after sublimation of the crude product under high vacuum at 70 °C (67 mg, 41% based on HBPin). mp 82-83°C. 1H NMR (CDC13) 5 1.32 (s, 12H, BOszle). 2.49 (s, 6H, 2Me), 7.28 (s, 2H). “B NMR (CDC13) 6 30.8. Anal. Calcd for C13H20BN02: C, 66.98; H, 8.65; N, 6.01. Found: C, 66.79; H, 9.09; N, 6.40. GC-MS (m/z) 233. quwaz BPin C6H4(C(O)NEtz)(BPin) (isomer mixture). Catalytic addition of HBPin to C6H5(C(O)NEt2) using solutions of 3 gave m-C6Ha(C(O)NEt2)(BPin) : p- C6H4(C(O)NEt2)(BPin) : 0-C6Ha(C(O)NEt2)(BPin) in the ratio 28: 14:58. The product was distilled from the reaction mixture using a Kugelrohr distillation apparatus and collected 151 as a colorless viscous liquid (106 mg, 50% based on HBPin). For isomer mixture: 1H NMR (CDCl3) 51.04 (m, 3H, CH3), 1.23 (m, 3H, CH3), 1.28, 1.33, 1.34 (s, 12H, B02C6H12), 3.20 (m, 2H, CH2), 3.55 (m, 2H, CH2), 7.23-7.26, 7.30-7.74, 7.75-7.81 (m, 4H). l3C(‘H} NMR (CDC13) 612.36, 12.78 (br), 13.56, 14.05 (br)(NCH2_(;H3), 24.76 (B02C2(QH3)4), 39.04 (br), 39.64, 42.87, 43.16 (br) (NQHZCH3), 83.23, 83.52, 83.82 (B02§_2(CH3)4), 125.27, 127.58, 128.08, 128.71, 130.34, 132.34, 134.64, 134.88, 135.16, 136.60, 139.74, 142.20 (aromatic resonances), 171.02, 171.17, 171.50 (carbonyl resonances). ”B NMR (CDC13) 6 28.7. IR (neat, cm") 660 (m), 783 (w), 858 (m), 965 (m), 1103 (m), 1146 (s), 1223 (w), 1287 (m), 1321 (m), 1356 (s), 1428 (m), 1634 (s). Satisfactory combustion analysis has not been obtained. GC-MS (m/z) 303. The identity of isomer products was established by comparing spectroscopic data to the independent synthesis of authentic samples using the literature method,25 and by comparing GC-MS data for the mixture to data for the independently prepared pure isomers. m- C6H4(C(O)NEt2)(BPin). 1H NMR (CDC13) 5 1.07 (t, 3H, CH3), 1.21 (t, 3H, CH3), 1.32 (s, 12H, BOzCsle), 3.22 (br, 2H, CH2), 3.52 (m, 2H, CH2), 7.30-7.47 (m, 3H), 7.79 (s, 1H). o-C6H4(C(O)NEt2)(BPin). 1H NMR (CDCl,) 61.04 (br, 3H, CH3), 1.28 (br, 3H, CH3), 1.28 (s, 12H, BOzCsle). 3.20 (q, 2H, CH2), 3.56 (q, 2H, CH2), 7.27 (d, 1H), 7.30-7.40 (m, 2H), 7.79 (d, 1H). p-C6H4(C(O)NEt2)(BPin). lH NMR (CDC13) 5 1.06 (br, 3H, CH3), 1.23 (br, 3H, CH3), 1.34 (s, 12H, BOngle), 3.19 (br, 2H, CH2), 3.53 (br, 2H, CH2), 7.34 (d, 2H), 7.81 (d, 2H). <\:\>-C(O)0Et BPin 152 C6H4(C(0)0Et)(BPin) (isomer mixture). Catalytic addition of HBPin to C6H5(C(O)0Et) using solutions of 3 gave m-C6H.t(C(O)OEt)(BPin) : p- CbHa(C(O)OEt)(BPin) : o-C6114(C(O)0Et)(BPin) in the ratio 57:33:10. The identity of isomer products were established by comparing GC-MS data for the mixture to data for the independently prepared pure isomers using the literature method.25 m- C6H4(C(O)0Et)(BPin). 1H NMR (CDCl3) 51.34 (s, 12H, BOzCsle), 1.38 (t, 3H, CH3), 4.36 (q, 2H, CH2), 7.42 (t, 1H), 7.96 (d, 1H), 8.11 (d, 1H), 8.44 (s, 1H). 11B NMR (CDC13) 6 30.8. p-C6H.t(C(O)OEt)(BPin). ‘H NMR (CDC13) 6 1.34 (s, 12H,B02C6H12), 1.38 (t, 3H, CH3), 4.36 (q, 2H, CH2), 7.84 (d, 2H), 8.00 (d, 2H). “B NMR (CDC13) 6 30.9. GC-MS (m/z) 276. F F‘QBPin F 2,4,6-C6H2F3(BPin). 3 (5 mg, 0.013 mmol) and HBPin (90 mg, 0.70 mmol) were dissolved in 550 11L of a 1:2 ratio of xylene-(110 and 1,3,5-C6H3F3, and heated at 150 °C for 30 min in a J. Young tube. The solution was then transferred to a vial and the solvent removed under vacuum at room temperature. The residue was chromatographed on a silica gel column, eluting with CHzClz, to yield 2,4,6-C6H2F3(BPin) (83 mg, 46% based on HBPin). 1H NMR (CDCl3) 6 1.34 (s, 12H, Bozcan), 6.58 (m, 2H). ”B NMR (CDC13) 6 29.4. ”17 NMR (CDCl3) 6—103.8(m, 21:), —97.1 (m, 1F). Anal. Calcd for C12H14BF302: C, 55.85; H, 5.47. Found: C, 56.08; H, 5.59. GC-MS (m/z) 258. The general procedure for the syntheses of the following arylboronate esters: 13 (2.9 mg, 0.007 mmol) and dmpe (1 mg, 0.007 mmol) dissolved in an arene (166 11L x 2) 153 were transferred to a J -Young NMR tube, which was charged with HBPin (51 11L, 0.351 mmol). Additional arene (166 11L) was used to wash the residue to the NMR tube. The reaction mixture was then heated at 150 °C for xx h. The reaction was monitored by the disappearance of the resonance of pinacolborane in the 11B NMR spectra. After the reaction was done, an aliquot was taken for GC—FID and GC-MS analyses. The reaction mixture was then transferred to a vial and the substrate was removed under vacuum (in some cases with gentle heating). The residue was dissolved in CHzClz and passed through a silica gel column using CHzClz as eluting solvent. The filtrate was then pumped down to give the corresponding spectroscopically pure arylboronate esters. PinBQ-Cl Cl 1,2,4-C6H3(Cl)2(BPin). The borylation product was isolated as colorless oil (93.9 mg, 98%). 1H NMR (CDC13) 6 1.32 (s, 12 H, Bozcan), 7.41 (d, J = 8.0 Hz, 1H), 7.57 (dd, J = 8.0 Hz, 1.5 Hz, 1H), 7.84 (d, J: 1.5 Hz, 1H). 13C NMR (125 MHz, CDC13) 6 24.8, 84.3, 130.0, 132.3, 133.8, 135.5, 136.6. “B NMR (CDCl,) 6 30.0. Anal. Calcd for CizHlsBClez: C, 52.80; H, 5.54. Found: C, 53.09; H, 5.77. GC-MS (m/z) 273. CI c1 BPin 1,4,5-C6H3(Cl)2(BPin). The borylation product was isolated as a colorless solid (72.6 mg, 76%). mp 46-48 °C. 1H NMR (CD2C12) 6 1.35 (s, 12H, BOzCole), 7.29 (d, J = 8.3 Hz, 1H), 7.33 (dd, J= 2.4 Hz, 8.8 Hz, 1H), 7.67 (d, J= 2.4 Hz, 1H). l3C(‘H} NMR (CDC13) 6 24.8, 84.5, 130.7, 131.7, 132.1, 136.0, 137.7. “B NMR (CDC13) 6 29.9. Anal. Calcd for CrzHrsBCIZOZ: C, 52.80; H, 5.54. Found: C, 53.19; H, 5.71. GC-MS (m/z) 273. 154 PinB—Q—Me Me 1,2,4-C6H3(Me)z(BPin). The borylation product was isolated as colorless oil (69.2 mg, 85%). 1H NMR (CDC13) 6 1.32 (s, 12H, 302C6H12), 2.25 (s, 3H, Me), 2.26 (s, 3H, Me), 7.12 (d, J: 7.3 Hz, 1H), 7.53 (d, J: 7.3 Hz, 1H), 7.57 (s, 1H). '3C{‘H} NMR (CDC13) 6 19.4, 20.0, 24.8, 83.5, 129.1, 132.4, 135.8, 135.9, 140.1. “B NMR (CDC13) 6 30.6. Anal. Calcd for C14H21B02: C, 72.44; H, 9.12. Found: C, 72.26; H, 8.98. GC-MS (m/z) 232. Me Me BPin 1,2,4-C6H3(BPin)(Me)2. The borylation product was isolated as colorless oil (55.4 mg, 68%). 1H NMR (CDC13) 6 1.33 (s, 12H, B02C6H12), 2.29 (s, 3H, Me), 2.48 (s, 3H, Me), 7.04 (d, J: 7.7 Hz, 1H), 7.11 (dd, J= 2.2 Hz, 7.7 Hz, 1H), 7.56 (d, J: 2.2 Hz, 1H). 13C(‘H} NMR (CDC13) 6 20.7, 21.7, 24.8, 83.3, 129.7, 131.5, 133.8, 136.3, 141.7. “B NMR (CDC13) 6 30.9. Anal. Calcd for CnHeroz: C, 72.44; H, 9.12. Found: C, F3C‘QCI A BPin 72.00; H, 8.59. GC-MS (m/z) 232. C6H3(Cl)(BPin)(CF3) (isomer mixture). The borylation product was isolated as colorless oil (375 mg, 78%). The ratio of 1,2,4-C6H3(Cl)(BPin)(CF3) to 1,3,4- C6H3(Cl)(BPin)(CF3), determined by GC-FID of the crude reaction mixture, was 88:12. 1,2,4-C6H3(Cl)(BPin)(CF3). lH NMR (CDC13) 5 1.36 (s, 12H, BOzCsH12)s 7.44 (d, J = 8.5 Hz, 1H), 7.56 (ddd, J: 8.3 Hz, 2.4 Hz, 0.7 Hz, 1H), 7.92 (d, 2.2 Hz, 1H). 13C(‘H} 155 ma (CDC13) 6 24.8, 84.7, 124.0 (q, J: 272.0 Hz, 1C), 128.4 (q, J: 3.5 Hz, 1C), 128.5 (q, J: 32.9 Hz, 1C), 129.9, 133.3 (q, J: 3.5 Hz, 1C), 143.5. ”B NMR (CDC13) 6 30.0. “’1: NMR (CDC13) 6 —62.8. Anal. Calcd for C13H15BC1F302: C, 50.94; H, 4.93. Found: C, MeO—C\>—F —\ BPin 51.27; H, 5.05. GC-MS (m/z) 306. C6H3(F)(0Me)(BPin) (isomer mixture). The borylation product (55 mg, 62%) was isolated as colorless oil. The ratio of 1,4,6-C6H3(F)(0Me)(BPin) to 1,4,5- C6H3(F)(0Me)(BPin), determined by GC-FID of the crude reaction mixture, was 93:7. 1,4,6-C6H3(F)(0Me)(BPin). 1H NMR (CDC13) 6 1.34 (s, 12H, 802C6H12), 3.78 (s, 3H, OMe), 6.92 (m, 2H), 7.18 (m, 1H). '3C{‘H} NMR (CDC13) 6 24.8, 55.8, 83.9, 116.0 (d, 2JCF = 25.9 Hz) 119.2 (d, 3Jcp = 8.6 Hz), 120.1 (d, 3Jcp = 8.6 Hz), 155.3 (d, 4JCF = 1.9 Hz), 161.7 (d, lJCF = 243.6 Hz). “B NMR (CDC13) 6 29.8. ”’1: NMR (CDC13) 6 —114.3. 1,4,5-C6H3(F)(0Me)(BPin). 1H NMR (CDC13) 6 1.30 (s, 12H,BOzC6H12), 3.88 (s, 3H, OMe), 6.88 (d, J = 8.2 Hz, 1H), 7.64 (dd, J= 8.2 Hz, 1.5 Hz, 1H), 7.77 (d, J: 1.5 Hz, 1H). l3C(‘H} NMR (CDC13) 6 24.8, 56.7, 83.7, 112.2 (d, 3JCF = 6.7 Hz) 118.3 (d, 2JCF = 23.0 Hz), 122.4 (d, 2JCF = 21.1 Hz), 157.0 (d, lJCF = 238.8 Hz), 160.4 (d, 4JCF = 1.9 Hz). “B NMR (CDC13) 6 29.8. "’17 NMR (CDC13) 6 —125.6. Anal. Calcd for C13HrgBFO3: C, 61.94; H, 7.20. Found: C, 61.83; H, 7.45. GC-MS (m/z) 252. MeO—C\>‘Cl BPin C6H3(Cl)(OMe)(BPin) (isomer mixture). The borylation product was isolated as colorless oil (58 mg, 62%). The ratio of 1,4,6-C6H3(Cl)(OMe)(BPin) to 1,4,5- 156 C6H3(Cl)(OMe)(BPin), determined by GC-F ID of the crude reaction mixture, was 32:68. 1,4,6-C6H3(Cl)(OMe)(BPin). ‘H NMR (CDC13) 6 1.35 (s, 12H,BOzC6H12). 3.77 (s, 3H, OMe), 6.85 (dd, J: 8.6 Hz, 3.1 Hz, 1H), 7.17 (d, J: 3.3 Hz, 1H), 7.22 (d, J = 8.6 Hz, 1H). ‘3C(‘H} NMR (CDC13) 6 24.7, 55.5, 84.1, 117.9, 120.9, 130.2, 130.9, 157.6. “B NMR (CDC13) 6 30.2. 1,4,5-C6H3(Cl)(OMe)(BPin). 1H NMR (CDC13) 6 1.33 (s, 12H, BOzCole), 3.78 (s, 3H, OMe), 6.76 (d, J = 8.8 Hz, 1H), 7.30 (dd, J = 8.8 Hz, 2.9 Hz, 1H), 7.59 (d, J = 2.7 Hz, 1H). l3C(‘H} NMR (CDC13) 6 24.7, 56.1, 83.7, 112.1, 125.4, 131.9, 136.0, 162.7. “B NMR (CDC13) 6 30.2. Anal. Calcd for C13H18BC103: C, 58.14; H, 6.76. Found: C, 58.20; H, 6.97. GC-MS (m/z) 268. CI PinBQOMe C6H3(Cl)(OMe)(BPin) (isomer mixture). The borylation product was isolated as colorless oil (69 mg, 73%). The ratio of 1,2,4-C6H3(Cl)(OMe)(BPin) to 1,2,5- C6H3(Cl)(OMe)(BPin), determined by GC-FID of the crude reaction mixture, was 51:49. 1,2,5-C,H,(Cl)(0Me)(BPin). 'H NMR (CDC13) 6 1.31 (s, 12H, BOzC6H12), 3.88 (s, 3H, OMe), 6.88 (d, J: 8.3 Hz, 1H), 7.65 (dd, J: 8.1 Hz, 1.5 Hz, 1H), 7.78 (d, J: 1.5 Hz, 1H). I3C{'H} NMR (CDC13) 6 24.8, 56.0, 83.8, 111.4, 122.3, 134.7, 136.5, 157.4. ”B NMR (CDC13) 6 30.2. 1,2,4-C6Hs(Cl)(OMe)(BPin). ‘H NMR (CDC13) 6 1.32 (s, 12H, BOszle), 3.91 (s, 3H, OMe), 7.31-7.35 (m, 3H). 13C(‘H} NMR (CDC13) 6 24.8, 56.1, 84.0, 117.7, 126.0, 127.9, 129.7, 154.6. “B NMR (CDC13) 6 30.2. Anal. Calcd for C13H18BCIO3: C, 58.14; H, 6.76. Found: C, 58.12; H, 6.84. GC-MS (m/z) 268. Me / \ PinB/_ OMe 157 C6H3(0Me)(Me)(BPin) (isomer mixture). The borylation product was isolated as colorless oil (67 mg, 77%). The ratio of 1,2,5-C6H3(0Me)(Me)(BPin) to 1,2,4- C6H3(0Me)(Me)(BPin), determined by GC-FID of the crude reaction mixture, was 64:36. 1,2,5-C,H3(0Me)(Me)(BPin). lH NMR (CDC13) 6 1.34 (s, 12H, BOszle), 2.24 (s, 3H, Me), 3.87 (s, 3H, OMe), 7.15 (d, J: 7.3 Hz, 1H), 7.23 (s, 1H), 7.33 (d, J= 7.1 Hz, 1H). 13C(‘H} NMR (CDC13) 6 16.4, 24.8, 55.4, 83.6, 115.5, 127.3, 130.2, 130.3, 157.4. ”B NMR (CDC13) 6 30.6. 1,2,4—C6H3(0Me)(Me)(BPin). 1H NMR (CDC13) 6 1.33 (s, 12H, B02C6H12), 2.22 (s, 3H, Me), 3.84 (s, 3H, OMe), 6.82 (d, J: 8.2 Hz, 1H), 7.59 (s, 1H), 7.65(d, J: 8.2 Hz, 1H). ‘3C{'H} NMR (CDC13) 6 15.9, 24.8, 55.2, 83.5, 109.3, 125.9, 134.3, 137.2, 160.5. ”B NMR (CDC13) 6 30.6. Anal. Calcd for C14H21B03: C, 67.77; H, 8.53. Found: C, 67.76; H, 8.39. GC-MS (m/z) 248. MeO / \ Me BPin C6H3(0Me)(Me)(BPin) (isomer mixture). The borylation product was isolated as colorless oil (67 mg, 77%). The ratio of 1,4,6-C6H3(0Me)(Me)(BPin) to 1,4,5- C6H3(0Me)(Me)(BPin), determined by GC-FID of the crude reaction mixture, was 70:30. l,4,6-C6H3(0Me)(Me)(BPin). 1H NMR (CDC13) 5 1.34 (s, 12H, BOzCole). 2.26 (s, 3H, Me), 3.78 (s, 3H, OMe), 6.74 (d, J = 8.4 Hz, 1H), 7.16 (ddd, J = 8.4 Hz, 2.4 Hz, 0.7 Hz, 1H), 7.45 (d, J = 2.4 Hz, 1H). 13C{'H} NMR (CDC13) 5 20.2, 24.8, 56.1, 83.3, 110.8, 129.2, 132.8, 137.0, 162.4. ”B NMR (CDC13) 6 30.7. 1,4,5-C6H3(0Me)(Me)(BPin). 1H NMR (CDC13) 5 1.32 (s, 12H, BOzCoth), 2.45 (s, 3H, Me), 3.78 (s, 3H, OMe), 6.85 (dd, J= 8.4 Hz, 3.1Hz, 1H), 7.05 (d, J= 8.2 Hz,1H), 7.28 (d, J= 2.9 Hz, 1H). 13C{1H} NMR (CDC13) 5 21.1, 24.8, 55.3 (d, J= 2.0 Hz), 83.4, 117.0, 120.3, 130.8, 136.8, 156.9. HB 158 NMR (CDC13) 5 30.7. Anal. Calcd for C14H21BO3: C, 67.77; H, 8.53. Found: C, 67.50; H, 8.61. GC-MS (m/z) 248. C6H3(Cl)(Me)(BPin) (isomer mixture). The borylation product was isolated as colorless oil (74 mg, 84%). The ratio of 1,4,5-C6H3(CI)(Me)(BPin) to 1,4,6— C6H3(Cl)(Me)(BPin), determined by GC-FID of the crude reaction mixture, was 57:43. 1,4,5-C6H3(Cl)(Me)(BPin). ‘H NMR (CDC13) 6 1.32 (s, 12H, BozczHrz), 2.47 (s, 3H, Me), 7.06 (d, J: 8.3 Hz, 1H), 7.24 (dd, J: 8.3 Hz, 2.4 Hz, 1H), 7.69 (d, J: 2.4 Hz, 1H). l3C(‘H} NMR (CDC13) 6 21.5, 24.9, 83.8, 130.5, 130.8, 131.2, 135.4, 143.1. ”B NMR (CDC13) 6 30.9. 1,4,6—C,H3(Cl)(Me)(BPin). ‘H NMR (CDC13) 6 1.35 (s, 12H, BOszle). 2.28 (s, 3H, Me), 7.11 (ddd, J: 8.3 Hz, 2.0 Hz, 0.7 Hz, 1H), 7.21 (d, J= 8.3 Hz, 1H), 7.46 (d, J = 2.0 Hz, 1H). '3C{'H} NMR (CDC13) 6 20.6, 24.8, 84.1, 129.2, 132.5, 135.4, 136.6, 136.9. “B NMR (CDC13) 6 30.9. Anal. Calcd for C13H13BC102: C, 61.83; H, 7.18. Found: C, 62.15; H, 7.22. GC-MS (m/z) 252. CI PinBL_\ Me C6H4(Cl)(Me)(BPin) (isomer mixture). The borylation products (79 mg, 89%) were isolated as colorless oil. The ratio of 1,2,5-C6H3(Cl)(Me)(BPin) to 1,2,4- C6H3(Cl)(Me)(BPin), determined by GC-FID of the crude reaction mixture, was 62:38. 1,2,5-C6H3(Cl)(Me)(BPin). 1H NMR (CDC13) 5 1.32 (s, 12H, B02C6H12). 2.38 (s, 3H, Me), 7.20 (d, J = 7.5 Hz, 1H), 7.52 (dd, J: 8.8 Hz, 0.9 Hz, 1H), 7.76 (s, 1H). 13C{"H} NMR (125 MHz, CDC13) 6 20.2, 24.9, 84.0, 130.5, 132.8, 134.3, 135.2, 139.2. ”B NMR 159 (CDC13) 6 30.9. 1,2,4-C6H3(Cl)(Me)(BPin). 1H NMR (CDC13) 6 1.33 (s, 12H, BOzCsle), 2.37 (s, 3H, Me), 7.32 (d, J: 8.0 Hz, 1H), 7.55 (dd, J: 7.5 Hz, 0.9 Hz, 1H), 7.65 (s, 1H). l3C(‘H} NMR (125 MHz, CDC13) 6 19.7, 24.9, 83.9, 128.6, 133.5, 135.3, 137.3, 137.8. ”B NMR (CDC13) 6 30.9. Anal. Calcd for C13H13BC102: C, 61.83; H, 7.18. Found: C, 61.94; H, 7.42. GC-MS (m/z) 252. Competitive Borylation Experiments Competition between Toluene and Anisole Using Solutions of 2 Cp*Ir(PMe3)H2 (2) (15 mg, 0.037 mmol) was weighed in a test tube, and a pre—mixed 1:1 mole ratio of toluene (256 mg) and anisole (300 mg) was added to dissolve the catalyst. Then HBPin (27 11L, 0.185 mmol) was added to the test tube via syringe. The solution was transferred to a J. Young NMR tube, and the reaction was heated at 150 °C in a constant temperature circulator. The conversion of the reaction was monitored by the disappearance of the resonance for pinacolborane in the 11B NMR spectrum. The isomer ratios were determined by integrating the peaks in the GC-MS spectra, after correcting for the response factors determined from equimolar mixtures of independently synthesized arylboronate esters. The ratio of 0-, m-, p-C6H4Me(BPin) : -, m-, p- C6H4(0Me)(BPin) is 38:62. Competition between Toluene and Benzotrifluoride Using Solutions of 2 The procedure is the same as that in the competition between toluene and anisole using solutions of 2. Toluene (242 mg) and benzotrifluoride (383 mg) were used. The ratio of 0-, m-, p-C6H4Me(BPin) : m-, p-C,H4(CF3)(BPin) is 11:89. 160 Competition between Cumene and N,N-Dimethylaniline Using Solutions of 2 The procedure is the same as that in the competition between toluene and anisole using solutions of 2. Cumene (280 mg) and N,N-dimethylaniline (282 mg) were used. The ratio of C6114CH(CH3)2(BPin) : o-, m-, p-C6Ha(NMe2)(BPin) is 69:31. Competition between Toluene and Anisole Using Solutions of 3 Cp*Rh(n4-C6Me6) (3) (5 mg, 0.013 mmol) was weighed in a test tube, and a pre-mixed 1: 1 mole ratio of toluene (256 mg) and anisole (300 mg) was added to dissolve the catalyst. Then HBPin (90 mg, 0.70 mmol) was added to the test tube via syringe. The solution was transferred to a J. Young NMR tube. The reaction was heated at 150 °C in a constant temperature circulator, and the reaction was judged to be complete by monitoring the disappearance of the resonance for pinacolborane in the 11B NMR spectrum. The isomer ratios were determined by integrating the peaks in the GC-MS spectra, after correcting for the response factors determined from equimolar mixtures of independently synthesized arylboronate esters. The ratio of o-, m-, p-C6H4Me(BPin) : -, m-, p- C6H4(0Me)(BPin) is 46:54. Competition between Toluene and Benzotrifluoride Using Solutions of 3 The procedure is the same as that in the competition between toluene and anisole using solutions of 3. Toluene (242 mg) and benzotrifluoride (383 mg) were used. The ratio of o-, m-, p-C6H4Me(BPin) : m-, p-C6H4(CF3)(BPin) is 27:73. Competition between Cumene and N,N-Dimethylaniline Using Solutions of 3 The procedure is the same as that in the competition between toluene and anisole using solutions of 3. Cumene (280 mg) and N,N-dimethylaniline (282 mg) were used. The ratio of C,H,CH(CH,)2(BPin) : 0-, m-, p-C,H.,(NMe2)(BPin) is 60:40. 161 Competition between Toluene and N,N-Dimethylaniline Using Solutions of 3 The procedure is the same as that in the competition between toluene and anisole using solutions of 3. Toluene (250 mg) and N,N-dimethylaniline (329 mg) were used. The ratio of o-, m-, p-CgHaMe(BPin): o-, m-, p-C6H4(NMe2)(BPin) is 59:41. Competition Experiments Using Solutions of 13 and dmpe The general procedure is illustrated by the competition reaction between m-xylene and 1,3-bis(trifluoromethyl)benzene using solutions of 13 and dmpe. (Ind)Ir(COD) (13) (2.9 mg, 0.007 mmol) and dmpe (1 mg, 0.007 mmol) were weighed into two separate GC vials, and a pre-mixed 1:1 molar ratio of m-xylene and 1,3-bis(trifluoromethyl)benzene solution (166 11L x 3) was added to dissolve the catalyst. The solution was transferred to a J. Young NMR tube, which was charged with HBPin (51 11L, 0.351 mmol) and the reaction mixture was heated at 150 °C in a constant temperature circulator. The conversion of the reaction was monitored by the disappearance of the resonance for pinacolborane in the HB NMR spectrum. The product ratios were determined by integrating the peaks in the GC-FID spectra. The ratio of 1,3,5-C6H3Me2(BPin) : 1,3,5- C6H3(CF3)2(BPin) is 35:96.5. The results of borylation of equimolar mixtures of two different substituted arenes catalyzed by 13 (2 mol%)/dmpe (2 mol%) are summarized in Table 13. Kinetics Experiments A typical experimental run for the reaction of Cp*Rh(PMeg)(Ph)(H) (4) with 12 equiv. of HBPin in C6D6 is described as follows: Two samples were prepared at the same 162 I?!“ Inumflw time. In a drybox, compound 4 (18 mg, 0.046 mmol) and HBPin (80 11L, 0.55 mmol) were placed in a 1 mL volumetric flask and the flask was filled with C6D6 to the mark. The solution in the volumetric flask was mixed well and distributed equally to two J. Young NMR tubes. The kinetic experiments were run twice at different temperatures to ensure the reproducibility. The temperature range is from 65 °C to 115 °C. The kinetics was carried out at 65, 75, 85, 95, 105, and 115 °C. The reactions were heated in a constant temperature oil bath (Cole-Palmer Polystat Constant Temperature Circulator). At specific intervals the NMR tubes were removed from the oil bath and quenched by rapid cooling in an ice bath. The 1H N1V£R spectra were then recorded at 25 i 0.5 °C on a VXR-SOO spectrometer. The progress of the reaction was monitored to 3 half-lives by measuring the ratio of the intensity of the Cp* of 4 versus the total “intensity” of the Cp* resonances of 4 and Cp*Rh(PMe3)(H)(BPin) (7). A typical experimental run for the phosphine dependence on the thermolysis of 18 in Cng is described as follows: Four samples were prepared at the same time. In a drybox, compound 18 (60 mg, 0.096 mmol) was placed in a 1 mL volumetric flask and the flask was filled with C6D6 to the mark. The solution in the volumetric flask was mixed well and a 200 uL portion of the solution was added to four J. Young NMR tubes respectively via an auto-pipette (100 11L x 2). C6Me6 (15.6 mg, 0.096 mmol) was placed in a 1 m1 volumetric flask and the flask was filled with C6D6 to the mark. The solution in the volumetric flask was mixed well and a 200 uL portion of the solution was added to the four J. Young NMR tubes respectively via an auto-pipette (100 11L x 2) as an internal standard. Pre-calculated amount of Cng, was added into each of the four NMR tubes to make the total volume of the solution to 700 1.1L. In the end, different quantities of PMe3 163 were added to the four NMR tubes respectively via a microsyringe. The experiments were carried out at 130 °C in a constant temperature oil bath (Cole-Palmer Polystat Constant Temperature Circulator). At specific intervals the NMR tubes were removed from the oil bath and quenched by rapid cooling in an ice bath. The 1H NMR spectra were then recorded at 25 :1: 0.5 °C on a Inova-600 spectrometer. The concentration range of PMe3 is from 0.00828 M to 0.828 M. The progress of the reaction was monitored to 3 half-lives by measuring the ratio of the intensity of the PMe3 of 18 versus the total “intensity” of the PMe3 resonances of 18 and 37-d1. Crystal Structure Determinations and Refinement Crystals grown at —30 °C were covered in Paratone N and moved quickly from a scintillation vial to a microscope side. A suitable crystal was chosen and mounted on a glass fiber. The data collection were carried out at a sample temperature of 173(2) K on a Bruker AXS three-circle goniometer with a CCD detector. The data were processed and reduced utilizing the program SAINT PLUS supplied by Bruker AXS. The structure were solved by direct methods (SHELXTL v5.1, Bruker AXS) in conjunction with standard difference Fourier techniques. The figures shown were produced using ORTEP and ellipsoids are at the 25% probability level. Tables of pertinent data collection parameters for all compounds crystallographically characterized are given in appendix A. Single crystals of 14 were grown from a pentane solution at —30 oC and the structure was further confirmed by single—crystal X-ray crystallographic analysis. Single crystals of 17 and B2Pin2 co-crystallized from a pentane solution at —30 °C and the structures were established by X-ray crystallographic analysis. 164 Single crystals of 18 were grown from a pentane solution at —30 °C and the structure was further confirmed by single—crystal X-ray crystallographic analysis. Crystals suitable for X-ray analysis of 25 were grown from a pentane solution at —30 °C. Single crystals of 28 were grown fiom a pentane solution at —30 0C to give colorless crystals suitable for X-ray analysis. Single crystals of 29 were grown from a concentrated pentane solution and the structure was further confirmed by single—crystal X-ray crystallographic analysis. 165 APPENDICES 166 Appendix A. Surmnary of crystal data and structure refinement for compound 14. Empirical formula Formula weight Temperature (K) Wavelength (A) Crystal system Space group a (A) b (A) C (A) 01, deg 13. deg 16 deg Volume (A3) Z Density (calculated) (Mg/m3) Absorption coefficient (mm'l) F (000) Crystal size (mm3) Theta range for data collection, deg Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>2sigma(I)] R indices (all data) Largest diff. peak and hole (en/3(3) (32711451331106 690.26 173(2) 0.71073 Monoclinic P2(1)/n 10.211(3) 16.822(4) 18.362(5) 90 92.907(5) 90 3150.2(14) 4 1.455 4.273 1388 0.32 x 0.30 x 0.28 1.64 to 23.28 -1 1<=h<=10, -18<=k<=18, -20<=l<=14 14167 4535 [R(int) = 0.1070] Full-matrix least-squares on F 2 4535 / 0 / 349 0.944 R1 = 0.0453, wR2 = 0.1056 R1= 0.0715, wR2 = 0.1146 1.604 and -1.635 167 Appendix A (cont). Summary of crystal data and structure refinement for compound 17. Empirical formula Formula weight Temperature (K) Wavelength (A) Crystal system Space group a (A) b (A) c (A) 01, deg B. deg 7, deg Volume (A3) Z Density (calculated) (Mg/m3) Absorption coefficient (mm'l) F(000) Crystal size (mm3) Theta range for data collection, deg Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>2sigma(I)] R indices (all data) Largest diff. peak and hole (e.A'3) C21H5132CIIIO4P3 557.84 173(2) 0.71073 Monoclinic P2(1)/c 11.4474(13) l7.2608(19) 19.579(2) 90 92.162(2) 90 3865.9(8) 6 1.438 3.681 1708 0.42 x 0.40 x 0.38 1.57 to 23.28 -12<=h<=12, -19<=k<=19, -21<=1<=21 31832 5555 [R(int) = 0.1432] Full-matrix least-squares on F 2 5555 / 0 / 391 1.052 R1 = 0.0320, wR2 = 0.0874 R1 = 0.0363, wR2 = 0.090 1.786 and -1.656 168 Appendix (cont). Summary of crystal data and structure refinement for compound 18. Empirical formula Formula weight Temperature (K) Wavelength (A) Crystal system Space group a (A) b (A) c (A) 0t, deg 13, deg 16 deg Volume (A3) Z Density (calculated) (Mg/m3) Absorption coefficient (mm'l) F(000) Crystal size (mm3) Theta range for data collection, deg Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>2$igma(I)] R indices (all data) Largest diff. peak and hole (e.A'3) C18H481311’02P4 31 1.73 173(2) 0.71073 Triclinic P-l 9.290(3) 12.408(5) 12.458(5) 90.107(5) 99.436(6) 90.221(6) 1416.4(9) 4 1.462 4.949 628 0.43 x 0.31 x 0.28 1.64 to 23.31 -10<=h<=7, -13<=k<=13, -12<=1<=13 6368 4039 [R(int) = 0.0226] Full-matrix least-squares on F 2 4039 / 0 / 252 1.103 R1 = 0.0380, wR2 = 0.1022 R1 = 0.0394, wR2 = 0.1035 2.484 and -1.546 169 Appendix (cont). Summary of crystal data and structure refinement for compound 25. Empirical formula Formula weight Temperature (K) Wavelength (A) Crystal system Space group a (A) b (A) c (A) 0t, deg B. deg 16 deg Volume (A3) Z Density (calculated) (Mg/m3) Absorption coefficient (mm'l) F(000) Crystal size (mm3) Theta range for data collection, deg Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>2sigma(I)] R indices (all data) Largest diff. peak and hole (e.A'3) C27H63B3II‘06P3 801.31 173(2) 0.71073 Monoclinic P2(l)/c 17.808(4) 11.073(3) 19.023(5) 90 90.106(4) 90 3751.0(15) 5 1.774 4.651 2050 0.44 x 0.28 x 0.26 2.13 to 23.28 -19<=h<=18, -12<=k<=12, -12<=1<=21 16587 5396 [R(int) = 0.0595] Full-matrix least-squares on F2 5396 / 0 / 362 1.037 R1 = 0.0743, wR2 = 0.1796 R1 = 0.1012, wR2 = 0.1971 3.058 and -1.864 170 Appendix (cont). Summary of crystal data and structure refinement for compound 28. Empirical formula Formula weight Temperature (K) Wavelength (A) Crystal system Space group a (A) b (A) c (A) 0t, deg 13. deg Y: deg Volume (A3) Z Density (calculated) (Mg/m3) Absorption coefficient (mm'l) F(000) Crystal size (mm3) Theta range for data collection, deg Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>23igma(I)] R indices (all data) Largest diff. peak and hole (e.A'3) C21H44BII'02P3 499.59 173(2) 0.71073 Monoclinic P2(1)/n 10.740(2) 16.687(3) 15.145(3) 90 90.069(4) 90 2714.3(10) 5 1.528 5.109 1252 0.42 x 0.35 x 0.24 1.82 to 23.28 -11<=h<=11, -17<=k<=18, -16<=1<=16 12074 3908 [R(int) = 0.0288] Full-matrix least-squares on F2 3908 / 0 / 266 1.119 R1 = 0.0282, wR2 = 0.0647 R1 = 0.0346, wR2 = 0.0669 1.110 and -0.912 171 Appendix A (cont). Summary of crystal data and structure refinement for compound 29. Empirical formula Formula weight Temperature (K) Wavelength (A) Crystal system Space group a (A) b (A) c (A) 0t, deg 15. deg 16 deg Volume (A3) 2 Density (calculated) (Mg/m3) Absorption coefficient (mm'l) F(000) Crystal size (mm3) Theta range for data collection, deg Index ranges Reflections collected Independent reflections Refinement method Data / restraints / parameters Goodness-of-fit on F2 Final R indices [I>2$igrna(I)] R indices (all data) Largest diff. peak and hole (e. A'3) C21H54B1r02P3Si 530.12 173(2) 0.71073 Monoclinic P2(l)/c l9.733(8) 10.280(4) 16.1 19(7) 90 1 10.168(7) 90 3069(2) 5 1.434 4.559 1348 0.36 x 0.34 x 0.26 2.20 to 23.28 -21<=h<=21, -7<=k<=l 1, —16<=l<=1 7 9405 4175 [R(int) = 0.1122] Full-matrix least-squares on F2 4175 / 0 / 278 1.056 R1 = 0.0472, wR2 = 0.1208 R1 = 0.0506, wR2 = 0.1238 3.122 and -2.075 172 ’31 1 g a. 1 ti APPENDIX B. Derivation of Rate Expressions for Chapter 5 k1 k [18]-————= (PMe3)3|r(BPin) + PMe3] 2 [(PMe3)2|r(BPin)+ PMe3] 1 [3] K2 191 R3 CSDS k4 C606 (PMe3)4lr(D) + ceogepin (PMe3)4|r(D) + CngBPin From the Figure above, the time dependent concentrations for 18, B, and C are governed by the Equations B1, B2, and B-3: d[18] _—dt_ = k1[18] ' k.1[B][PMe3] (Bl) d[B dt = k11131- KtlBlIPMea] - k213] + K2[C][L] - kBIBHCoDol (32) d C] —d, = kzlal - KzlcllLl - memos] (33) Application of the steady state approximation to [C] yields Equation B4, and combination of the steady state expression for [B] and Equation B4 gives B5. d[C] _ k213] dt ' 0 :3 [°] K2[PM931+ moons] 173 { it ~ -— (B4) ——d[B] k [18] ’ k [B][PMe ] + k [B] + k [B][C D ] k-ZKZIBHL] dt = 0 Z) 1 " " 3 2 3 6 6 K21PM631+ 1619st (BS) Rearrangement of Equation B5 gives Equation B6. The first order rate law in Equation B7 follows fi'om substitution of the expression for [B] (Equation B6) in Equation B1: B _ < k1ik-21PM33] + k4ICGDGD > [18] k—tk-le’lV'ea]2 + (k-1K4ICSDB] "’ szalcesDemPMeal + k4IC606Kk2 "‘ k3ICSDGD (B6) _ d[18] _ < k1K2k3IC606HPMeal + ktk4[0606](k2 + kalCeDc—sl) > 18] dt K1K21PM9312 + (K1k4106061 + K2k310606])[PMeal + k4lCeDol(k2 + k3[Cng]) (B7) From Equation B7, kobs (Equation B8) and l/kob, (Equation B9) can be extracted: k1K2k3[Cng][PM63] + k1k4[CGDB](k2 1’ k3lCBDGD kobs = k,1k-2[PMe3]2 + (K1k4 + k-2k3)[CSD6][PMe3] + MCGDGK'Q + k3lceDel) (BS) 174 1 1 k-lk-21PM9312 + k_1k4[Cng][PMe3] kobs k1 k-110606KK-2k31PMes] + k4k2 + k4kalceDsl) (39) If the reaction only goes through intermediate B to yield products (k4 = 0), l/kobS can be derived in Equation B10. If the reaction only passes through intermediate C to yield products (k3 = 0), l/kobs can be derived in Equation B11. 1 1 K1 — + k... ' k. ktkalCoDol [PMe3] 1 1 + K1K21PM9312+K1k41C606HPM33] kobs k1 k1k2k41C606] 175 (1310) (1311) APPENDIX C. Kinetic Details Data for the reaction of Cp*Rh(PMe3)(Ph)(H) (4) with 12 equiv. HBPin in C6D6. The progress of the reaction was monitored to 3 half-lives by measuring the ratio of the intensity of the Cp* of 4 versus the total “intensity” of the Cp* resonances of 4 and 7. The experiments were performed at various temperatures in a constant temperature oil bath. [4] = 0.046 M; [HBPin] = 0.551 M Temperature (°C) kobs (599.1) 65 1.27 x 10'5 75 4.0 x 10‘5 85 1.38 x 10'4 95 3.37 x 10‘4 105 9.8 x 10'4 115 2.0 x 10'3 Data for the reaction of Cp*Rh(PMe3)(Ph)(H) (4) with 24 equiv. HBPin in C6D6. The progress of the reaction was monitored to 3 half-lives by measuring the ratio of the intensity of the Cp* of 4 versus the total “intensity” of the Cp* resonances of 4 and 7. The experiments were performed at 95 i 0.5 °C in a constant temperature oil bath. 176 [HBPin] k,,,, (scc") 0.551 M 3.37 x 10“ 1.103 M 6.80x10“1 Data for the phosphine dependence on the thermolysis of (PMe3)4Ir(BPin) (18) in C6D6 at 130 °C. The progress of the reaction was monitored to 3 half-lives by measuring the ratio of the intensity of the PMe3 of 18 versus the total “intensity” of the PMe3 resonances of 18 and 37-d]. The experiments were performed at 130 :t 0.5 °C in a constant temperature oil bath. [18] = 0.0275 M; [C6Me6] = 0.0275 M [PMesl (M) k... (sec") 0 0.0417 0.0083 0.0170 0.0138 0.0138 0.0207 0.0097 0.0276 0.0072 0.0552 0.0044 0.1380 0.0021 177 BIBLIOGRAPHY 178 BIBLIOGRAPHY (l) (a) Bergman R. G. Science 1984, 223, 902-908. (b) Jones, W. D.; Feher, F. J. Acc. Chem. Res. 1989, 22, 91-100. (c) Amdtsen, B. A.; Bergman, R. G.; Mobley, T. A.; Peterson, T. H. Acc. Chem. Res. 1995, 28, 154-162. ((1) Crabtree R. H. Chem. Rev. 1995, 95, 987-1007. (e) Bengali, A. A.; Amdtsen, B. A.; Burger, P. M.; Schulta, R. H.; Weiller, B. H.; Kyle, K. R.; Moore, C. B.; Bergman, R. G. Pure Appl. Chem. 1995, 67, 281-288. (0 Stahl, S. S.; Labinger, J. A.; Bercaw, J. E. Angew. Chem, Int. Ed. Engl. 1998, 37, 2180-2192. (g) Crabtree R. H. J. Chem. Soc., Dalton Trans. 2001, 2437-2450. (h) Labinger J. A.; Bercaw, J. E. Nature 2002, 41 7, 507-514. (2) Olah, G. A.; Schleyer, P. v. R.;, Eds. “Carbonium Ions”; Wiley: New York, Vol. 1, 1968; Vol. 2, 1970; Vol. 3, 1972; Vol. 4, 1973; Vol. 5, 1976. (3) Olah, G. A.; Yoneda, N.; Parker, D. G. J. Am. Chem. Soc. 1976, 98, 5261-5268. (4) (a) Selective Hydrocarbon Activation (Eds: Davies, J. A.; Watson, P. L..; Liebman, J. F.; Greenberg, A.), VCH, New York, 1990. (b) Shilov, A. E. Activation of Saturated Hydrocarbons by Transition Metal Complexes, Reidel, Dordrecht, 1984. (c) Labinger, J. A., Fuel Process, Technol. 1995, 42, 325-338. ((1) Shilov, A. E.; Shul’pin Chem. Rev. 1997, 97, 2879-2932. (5) (a) F arinas, E. T.; Schwaneberg, U.; Glieder, A.; Arnold, F. H. Advanced Synthesis & Catalysis 2001, 343, 601-606. (b) Costas, M.; Chen, K.; Que, L., Jr. Coord. Chem. Rev. 2000, 200-202, 517-544. (c) Mansuy, D.; Battioni, P. Act. F unct. Alkanes 1989, 195-218. ((1) Atkins, W. M.; Sligar, S. G. J. Am. Chem. Soc. 1989, 111, 2715-2717. (6) (a) Neimann, K.; Neumann, R.; Rabion, A.; Buchanan, R. M.; Fish, R. H. Inorg. Chem. 1999, 38, 3575-3580. (b) Rabion, A.; Chen, D. S.; Wang, J.; Buchanan, R. M.; Seris, J .-L.; Fish, R. H. J. Am. Chem. Soc. 1995, 117, 12356-12357. (7) Kitajima, N.; Fukui, H.; Moro-oka, Y. J. Chem. Soc., Chem. Commun. 1988, 485. (8) Janowicz, A. H.; Bergman, R. G. J. Am. Chem. Soc. 1982, 104, 352-354. (9) (a) Watson, T. L. J. Am. Chem. Soc. 1983, 105, 6491-6493. (b) Thompson, M. E.; Baxter, S. M.; Bulls, A. R.; Burger, B. J .; Nolan, M. C.; Santarsiero, B. D.; Schaefer, W. P.; Bercaw, J. E. J. Am. Chem. Soc. 1987, 109, 203-219. (10) (a) Cummins, C. C.; Baxter, S. M.; Wolczanski, P. T. J. Am. Chem. Soc. 1988, 110, 8731-8733. (b) Walsh, P. J.; Hollander, F. J.; Bergman, R. G. J. Am. Chem. Soc. 1988, 110, 8729-8731. (c) Bennett, J. L.; Wolczanski, P. T. J. Am. Chem. Soc. 1997, 119, 179 10696-10719. ((1) Schaefer, W. P.; Wolczanski, P. T. J. Am. Chem. Soc. 1998, 120, 4881- 4882. (11) (a) Sherry, A. E.; Wayland, B. B. J. Am. Chem. Soc. 1990, 112, 1259-1261. (b) Wayland, B.B.; Sherry, A. E. J. Am. Chem. Soc. 1991, 113, 5305-5311. (12) Crabtree, R. H.; Mihelcic, J. M.; Quirk, J. M. J. Am. Chem. Soc. 1979, 101, 7738- 7739. (13) (a) Fisher, B. J .; Eisenberg, R. Organometallics 1983, 2, 764-767. (b) Sakakura, T.; Sodeyarna, T.; Sasaki, K.; Wada, K.; Tanaka, M. J. Am. Chem. Soc. 1990, 112, 7221- 7229. (14) Sakakura, T.; Tanaka, M. J. Chem. Soc., Chem. Commun. 1987, 10, 758-759. (15) (a) Jenson, C. M. J. Chem. Soc., Chem. Commun. 1999, 2443-2449. (b) Liu, F. C.; Pak, E. B.; Singh, B.; Jensen, C. M.; Goldman, A. S. J. Am. Chem. Soc. 1999, 121, 4086- 4087. (16) (a) Rablen, P. R.; Hartwig, J. F. J. Am. Chem. Soc. 1994, 116, 4121-4122. (b) Rablen, P. R.; Hartwig, J. F. J. Am. Chem. Soc. 1996, 118, 4648-4653. (17) Waltz, K. M.; He, X.; Muhoro, C.; Hartwig, J. F. J. Am. Chem. Soc. 1995, 117, 11357-11358. (18) (a) Westcott, S. A.; Blom, H. P.; Marder, T. E; Baker, R. T. J. Am. Chem. Soc. 1992, 114, 8863-8869. (b) Burgess, K.; van der Donk, W. A.; Westcott, S. A.; Marder, T. R; Baker, R. T. Calabrese, G. C. J. Am. Chem. Soc. 1992, 114, 9350-9359. (c) Brown, J, M.; Lloyd-Jones, G. C. J. Chem. Soc., Chem. Commun. 1993, 710. (19) Waltz, K. M.; Hartwig, J. F. Science 1997, 277, 211-213. (20) Waltz, K. M.; Muhoro, C. N.; Hartwig, J. F. Organometallics 1999, 18, 3383-3393. (21) Iverson, C. N.; Smith M. R. III J. Am. Chem. Soc. 1999, 121, 7696-7697. (22) (a) Periana, R. A.; Taube, D. J .; Evitt, E. R.; Loffler, D. G.; Wentrcek, P. R.; Voss, G.; Masuda, T. Science 1993, 259, 340-343. (b) Periana, R. A.; Taube, D. J .; Gamble, S.; Taube, H.; Satoh, T.; Fujii, H. Science 1998, 280, 560-564. (23) (a) Miyaura, N.; Suzuki, A. Chem. Rev. 1995, 95, 2457-2483. (b) Suzuki, A. J. Organomet. Chem. 1999, 576, 147-168. (24) Chen, H; Hartwig, J. F. Angew. Chem. Int. Ed. Engl. 1999, 38, 3391-3393. 180 (25) Ishiyama, T.; Murata, M.; Miyaura, N. J. Org. Chem. 1995, 60, 7508-7510. (26) (a) Chen, H; Schlecht, S.; Semple, T. C.; Hartwig, J. F. Science, 2000 287, 1995- 1997. (b) Cho, J.-Y., Iverson, C. N.; Smith, M. R. 111 J. Am. Chem. Soc. 2000 122, 12868-12869. (c) Tse, M. K.; Cho, J.-Y.; Smith M. R. III Org. Lett. 2001 3, 2831-2833. ((1) Shimada, S.; Batsanov, A. 8.; Howard, J. A. K.; Marder, T. B. Angew. Chem. Int. Ed. Engl. 2001, 40, 2168-2171. (e) Cho, J.-Y.; Tse, M. K.; Holmes, D.; Maleczka, R. E. Jr.; Smith, M. R. 111 Science 2002, 295, 305-308. (0 Ishiyama, T.; Takagi, J.; Ishida, K.; Miyaura, N.; Anastasi, N. R.; Hartwig, J. F. J. Am. Chem. Soc. 2002 124, 390-391. (g) Takagi, J .; Kazuaki, S.; Hartwig, J. F.; Ishiyama, T.; Miyaura, N. Tetrahedron Lett. 2002 43, 5649-5651. (27) Goldfuss, B.; Knochel, P.; Bromm, L. 0.; Knapp K. Angew. Chem. Int. Ed. Eng]. 2000, 39, 4136-4139. (28) Aizenberg, M.; Milstein, D. Science 1994, 265, 359-361. (29) Jones, W. D.; Feher, F. J. J. Am. Chem. Soc. 1984, 106, 1650-1663. (30) Kawamura, K.; Hartwig, J. F. J. Am. Chem. Soc. 2001, 123, 8422-8423. (31) Nguyen, P.; Blom, H. P.; Westcott, S. A.; Taylor, N. J .; Marder, T. B. J. Am. Chem. Soc. 1993, 115, 9329-9330. (32) The approach was first demonstrated by Dr. Man Kin Tse. (33) Some of the experiments were carried out by Dr. Man Kin Tse. (34) Anastasi, N. R.; Hartwig, J. F. J. Am. Chem. Soc. 2002, 124, 390-391. (35) Ezbiansky, K.; Djurovich, P. I.; LaForest, M.; Sinning, D. J .; Zayes, R.; Berry, D. H. Organometallics 1998, I 7, 1455-1457. (36) Murata, M.; Oyama, T.; Watanabe, S.; Masuda, Y. J. Org. Chem. 2000, 65, 164-168. (37) White, D. P.; Anthony, J. C.; Oyefeso, A. O. J. Org. Chem. 1999, 64, 7707-7716. (38) Winstein, S.; Holness, N. J. Am. Chem. Soc. 1955, 77, 5562-5578. (39) CRC Handbook of Chemistry and Physics, 831rd Edition 2002, Lide, D. L. CRC Press. (40) (a) Dawans, F .; Morel, D. J. Mol. Catal. 1977-78, 3, 403. (b) Halpem, J .; Okamoto, T.; Zakhariev, A. J. Mol. Catal. 1976, 2, 65. (c) Chan, A. S. C.; Halpem, J. J. Am. Chem. 181 Soc. 1980, 102, 838. (d) Halpem, J. Inorg. Chim. Acta. 1981, 50, 11. (e) Halpem, J .; Riley, D. P.; Chan, A. C. S.; Pluth, J. J. J. Am. Chem. Soc. 1977, 99, 8055. (f) Halpem, J. Science 1982, 21 7, 401. (41) Harper, T. G. P.; Desrosiers, P. J .; Flood, T. C. Organometallics 1990, 9, 2523-2528. (42) Knorr, J. R.; Merola, J. S. Organometallics 1990, 9, 3008-3010. (43) Dai, C; Stringer, G; Marder, T. B.; Scott, A. J.; Clegg, W; Norman, N. C. Inorg. Chem. 1997, 36, 272-273. (44) Thorn, D. L.; Tulip, T. H. Organometallics 1982, 1, 1580-1586. (45) Harris, R. K. In Nuclear Magnetic Resonance Spectroscopy. Longman Scientific and Technical, Essex, UK. 1986 (46) Dai, C.; Stringer, G.; Marder, T. B.; Baker, R. T.; Scott, A. J .; Clegg, W.; Norman, N. C. Can. J. Chem. 1996, 74, 2026-2031. (47) Cleary, B.; Eisenberg, R. Organometallics 1995, 14, 4525-4534. (48) (a) Johnson, C. E.; Eisenberg, R. J. Am. Chem. Soc. 1985, 107, 3148-3160. (b) Johnson, C. E.; Eisenberg, R. J. Am. Chem. Soc. 1985, 107, 6531-6540. (49) Sargent, A. L.; Hall, M. B.; Guest, M. F. J. Am. Chem. Soc. 1992, 114, 517-522. (50) Lu, Z.; Jun, C.-H.; de Gala, S. R.; Sigalas, M.; Eisenstein, O; Crabtree, R. H. J. Chem. Soc., Chem. Commun. 1993, 1877-1880. (51) Rickard, C. E. F.; Roper, W. R.; Williamson, A.; Wright, L. J. Angew. Chem. Int. Ed. 1999, 38,1110-1113. (52) Thorn, D. L. Organometallics 1982, 1, 197-204. (53) The ratio was determined from the integrations of pinacol resonance in the 1H NMR spectrum. (54) Baker, R. T.; Ovenall, D. W.; Calabrese, J. C.; Westcott, S. A.; Taylor, N. J.; Williams, I. D.; Marder, T. B. J. Am. Chem. Soc. 1990, 112, 9399-9400. (55) Aizenberg, M.; Milstein, D. J. Am. Chem. Soc. 1995, 117, 6456-6464. (56) Sakaki S.; Satoru, K.; Manabu, S. Organometallics 1999, 18, 4825-4837. 182 (57) See Table 4 in the Chapter 3. (58) Jones, W. D.; Feher, F. J. J. Am. Chem. Soc. 1986, 108, 4814-4819. (59) The product distribution between C6D5BPin and C6H5BPin was determined by integrating the peaks in the GC-FID spectra, after correcting for the response factors determined from equimolar mixtures of independently synthesized arylboronate esters. The product distribution between C6D2H3(BPin) and C6D3H2(BPin) was extracted from IH NMR spectrum of the crude mixture. T1 measurement experiments were conducted before acquiring 1H NMR spectra. (60) Ishiyama, T.; Takagi, J.; Hartwig, J. F.; Miyaura, N. Angew. Chem. Int. Ed. 2002, 41, 3056-3058. (61) Armarego, W. L. F.; Perrin, D. D. Purification of Laboratory Chemicals, Fourth Edition, Butterworth Heinemann. (62) Horde, J. L.; Lambert, J. C.; Senoff, C. V. Inorg. Synth. 1974, 15, 18-20. (63) Merola, J. S.; Kacmarcik, R. T. Organometallics 1989, 8, 778-784. (64) Herskovitz, T. Inorg. Synth. 1982, 21, 99-102. (65) Bowyer, W. J .; Merkert, J. W.; Geiger, W. E. Organometallics 1989, 8, 191- 198. (66) Bowyer, W. J.; Geiger, W. E. J. Am. Chem. Soc. 1985, 107, 5657-5663. (67) Janowicz, A. H.; Bergman, R. G. J. Am. Chem. Soc. 1983, 105, 3929-3939. (68) Barefield, E. K. Inorg. Synth. 1974, 15, 34-38. (69) Gilbert, T. M.; Bergman, R. G. Organometallics 1983, 2, 1458-1460. 183 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIII 1111[11]][11[Mlljljtltlll