...:..}.a, J... .1...‘... ‘ .rr 0.}. 1.2.1.: . ., )1 I .2. .. . .. .. ~ . 3):}. t:..-L...\ . WESW‘ llHIIIIUIHHIIWI”WilliIIUIHHWIHIHIHIIHNUI 300910 1829 This is to certify that the dissertation entitled Organometallic, Coordination and Redox Chemistry of Rhodium(II) Metalloradical Species Supported by an Oxygen Functionalized Triaryl Phosphine presented by Steven Christopher Haefner has been accepted towards fulfillment of the requirements for Ph.D. Chemistry degree in TWQWL. Major professor Date August 18, 1992 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 rWuemmv Michigan State University —_—— PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE MSU Is An Affirmative Action/Equal Opportunity Institution c:\clrc\datedue.pm&p.1 ORGANOMETALLIC, COORDINATION AND REDOX CHEMISTRY OF RHODIUM(II) METALLORADICAL SPECIES SUPPORTED BY AN OXYGEN FUNCTIONALIZED TRIARYL PHOSPHINE By Steven Christopher Haefner A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1992 ABSTRACT ORGANOMETALLIC, COORDINATION AND REDOX CHEMISTRY OF RHODIUM(II) METALLORADICAL SPECIES SUPPORTED BY AN OXYGEN FUNCTIONALIZED TRIARYL PHOSPHINE By Steven Christopher Haefner The study of mononuclear rhodium chemistry has focused primarily on its monovalent and trivalent oxidation states with relatively little emphasis on the chemistry of divalent species. The scarcity of mononuclear Rh(II) complexes is due, in part, to the proclivity of these systems to either undergo dimerization or disproportionation, consequently few paramagnetic Rh(II) complexes have been the subject of comprehensive studies. Recent reports of carbon monoxide and methane activation by Rh(II) metalloradicals have sparked renewed interest in this under developed area of chemistry. This work focuses on Rh(II) radical chemistry of the multifunctionalized ether phosphine ligand tris(2,4,6-trimethoxyphenyl)- phosphine (TMPP). The unique combination of steric bulk and chelating abilities afforded by TMPP has allowed for the stabilization of a novel mononuclear, six coordinate Rh(II) homoleptic phosphine complex, [Rh(n3- TMPP)2][BF4]2. The steric bulk of the ligand precludes the formation of dinuclear species, yet the presence of labile ether interactions permits the complex to react with a variety of small molecules. Most notably, this — i-_ A metalloradical species readily reacts with n-acceptors such as carbon monoxide and isocyanide ligands to form adducts of the type [Rh(TMPP)2L2]2+. In the case of CO, the complex is highly unstable and immediately undergoes a series of redox reactions involving the formation of Rh(I) and Rh(III) intermediates that ultimately regenerate the original Rh(II) complex. In contrast, the reaction of [Rh(n3-TMPP)2][BF4]2 with the weaker n-acceptor ligands CNR (R = tBu, iPr) affords the stable four-coordinate Rh(II) adduct [Rh(TMPP)2(CNR)2][BF4]2 . In addition to the observed substitution chemistry with n-acceptors, [Rh(n3-TMPP)2][BF4]2 reacts with nucleophiles, resulting in dealkylation of a bound methoxy-group to give a new Rh(II) complex ligated by one phosphine and one phosphino-phenoxide ligand. This complex, formulated as [Rh(TMPP)(TMPP-O)][BF4], (TMPP-0 = [P{C6H2(0Me)3}2(CGH2(OMe)2O)]1'), reacts with carbon monoxide to form unusual paramagnetic adducts that have been detected by IR and EPR spectroscopies. The series of d7 phosphine complexes isolated in these studies are among the first mononuclear Rh(II) complexes to be fully characterized by X- ray crystallography, EPR and a variety of other spectroscopic techniques. An account of the syntheses, characterization and reactivity of these rare paramagnetic species is presented together with preliminary results regarding the extension of this work to other radical systems. To my wife Kelly, without whose love and patience, this work would not have been possible. "What is essential is invisible to the eyes. It is only with the heart with which one can see rightly. " Antoine de Saint Expuréy ACKNOWLEDGEMENTS The research presented invthis dissertation would not have been possible without the assistance and support of numerous individuals whose direct and indirect contributions are an integral part of this work. Foremost, I would like to thank my research adviser, Professor Kim R. Dunbar. She taught me through example, the importance of hard work and dedication in achieving your goals. Her enthusiasm and love of chemistry encouraged me to reach within myself and strive for excellence. Above all, she impressed upon me to always be critical of oneself and work. I will always hold her advice and insight in the highest regard, as I have the deepest respect and admiration for Professor Dunbar as both a scientist and a friend. I extend my gratitude to my contemporaries in the Dunbar Group for sharing the successes and failures that go along with being a graduate student. Together, they made being a member of the Dunbar Group a memorable and enjoyable experience. In particular, I would like to thank Laura Pence for her help and friendship, especially in the early years when I was only a young pup still wet behind the ears. She impressed upon me the importance of promoting oneself and chemistry. Special thanks go to Anne Quillevéré; the other half of "the match made in heaven". Her support and encouragement went along way in helping me to achieve my goals. Although she would never admit to it, she truly shares my enthusiasm for chemistry. I will forever value and treasure — her friendship. I also thank my long time friends Gene, Jim and Rob for providing that often needed outlet when things did not work quite they way they should. Furthermore, I thank my parents for their love, support and for never expecting anything less from me. In addition, I would like to acknowledge Michigan State University, College of Natural Science, and Dow Chemical for providing financial support in the form of scholarships. Finally, I thank the numerous collaborators whom I have worked with over the past several years for their individual contributions to this work. TABLE OF CONTENTS _ Page LIST OF TABLES xvi LIST OF FIGURES xviii LIST OF SYMBOLS AND ABBREVIATIONS xxii LIST OF COMPOUNDS xxv CHAPTER I. INTRODUCTION 1 A. Role of Tertiary Phosphines in Transition Metal Chemistry ........ 2 B. Development of Ether-Phosphines 7 C. Chemical Aspects of Tris(2,4,6-trimethoxyphenyl)phosphine ....... 11 D. Chemistry of Mononuclear Rh(II) Complexes ................................ 18 List of References 24 CHAPTER II. STRUCTURAL AND SPECTROSCOPIC PROPERTIES OF TRIS(2,4,6- TRIMETHOXYPHENYL)PHOSPHINE AND TRIS(2,4,6-TRIMETHOXYPHENYL)PHOSPHINE OXIDE 32 1_ Intrnrlnofinn 33 2. Experimental 34 A. Synthesis 34 (1) Preparation of Tris(2,4,6- trimethoxyphenyl)phosphine (TMPP) ...................... 34 (2) Preparation of Tris(2,4,6- trimethoxyphenyl)phosphine Oxide (TMPP=O) ...... 35 (3) Preparation of Ni(CO)3TMPP ................................... 36 B. X-ray Crystallography 36 (1) TMPP 36 (i) Data Collection and Reduction 36 vii (ii) Structure Solution and Refinement. .............. 38 (2) TMPP=O - 2 H20 39 (i) Data Collection and Reduction 39 (ii) Structure Solution and Refinement ............... 39 3. Results and Discussion 40 A. Synthesis and Characterization 40 B. Molecular Structures 44 (l) TMPP 44 (2) TMPP=O - 2 H20 49 List of References 54 i CHAPTER III. ISOLATION, CHARACTERIZATION AND REDOX CHEMISTRY OF THE MONONUCLEAR Rh( II) COMPLEX [Rh(n3-TMPP)2][BF4]2 ................................. 56 1. Intrndrwfinn 57 2. Experimental 58 A. Synthesis 58 (1) Preparation of [Rh(n3-TMPP)2][BF4]2 (3) ................. 58 (2) Preparation of [Rh(n3-TMPP)2][BF4]3 (4) ................. 59 (i) Oxidation of [Rh(n3-TMPP)2][BF4]2 with NOBF4. 59 (ii) Oxidation of [Rh(n3-TMPP)2][BF4]2 with [CpZFe][BF4]. 60 (iii) Oxidation of [Rh(n3-TMPP)2][BF4]2 with HBF4 . Et20 60 (3) Demethylation of [Rh(n3-TMPP)2][BF4]3: Formation of [Rh(n3-TMPP)(CBH2(OMe)2- OP{C6H2(OMe)3}2)][BF4]2 (5) 60 (i) Solution 60 (ii) Solid State 61 (iii) Reaction of [Rh(n3-TMPP)2][BF4]3 (4) with TMPP 61 (iv) Reaction of [Rh(n3-TMPP)2][BF4]3 (4) with [(Bu”)4N][I] 61 (4) Chemical Reduction of [Rh(n3-TMPP)2][BF4]2 ....... 62 (5) Preparation of [Rh(TMPP)2][PF6]2 ........................... 62 viii B. X-ray Crystallography ........................................................... 63 (1) [Rh(n3-TMPP)2][BF4]2 (3) ......................................... 63 (i) Data Collection and Reduction ..................... 63 (ii) Structure Solution and Refinement. ............ 65 (2) [Rh(n3-TMPP)2][PF6]2[BF4] ..................................... .65 (i) Data Collection and Reduction. .................... 65 (ii) Structure Solution and Refinement. ............ 66 3. Results ............................................................................................... 67 A. Preparation and Spectroscopic Properties of [Rh(n3- TMPP)2][BF4]2 (3) ................................................................ 68 B. Magnetic and EPR Spectroscopic Properties of [Rh(n3- TMPP)2][BF4]2 (3) ................................................................ 7 1 C. Crystal Structure of [Rh(n3-TMPP)2][BF4]2 (3) ................. 75 D. Redox properties of [Rh(n3-TMPP)2][BF4]2 (3) ................... 80 (1) Oxidation of [Rh(n3-TMPP)2][BF4]2 (3) ................... 80 (i) Synthesis and Spectroscopic characterization of [Rh(n3-TMPP)2]- [BF4]3 ............................................................ 80 (ii) Crystal Structure of [Rh(n3-TMPP)2]- [PF6]2[BF4] .................................................... 84 (iii) Decomposition of [Rh(n3-TMPP)2][BF4]3 (4) ................................................................... 84 (2) Reduction of [Rh(n3-TMPP)2][BF4]2 (3) ................... 92 4. Discussion .......................................................................................... 94 List of References ................................................................................... 96 CHAPTER IV REVERSIBLE CARBON MONOXIDE CHEMISTRY OF [Rh(n3-TMPP)2][BF4]2 ...................... 100 1. Introduction ....................................................................................... 101 2. Experimental .................................................................................... 102 A. Synthesis ................................................................................ 102 (1) Reaction of [Rh(n3-TMPP)2][BF4]2 (3) with CO. ..... 103 (2) Reaction of [Rh(n3-TMPP)2][BF4]2 (3) with 1200/1300 (1:1). ....................................................... 103 (3) Preparation of [Rh(n2-TMPP)(TMPP)CO][BF4] (7 ) ................................................................................ 104 (i) Reduction of [Rh(n3-TMPP)2][BF4]2 1n the Presence of CO. ....................................... 104 (ii) Reaction of [Rh(cod)Cl]2 with TMPP in the Presence of CO. ....................................... 104 (iii) Reaction of [Rh(CO)2Cl]2 with TMPP ......... 105 (4) Preparation of [Rh(TMPP)2(CO)2][BF4] (6) ............ 105 (5) Solid State Reactions of (3) - (7) with CO. ................ 106 (6) Redox Reaction of [Rh(nZ-TMPP)(TMPP)CO]- [BF4] (7) with [Rh(n3-TMPP)2][BF4]3 (4). .............. .106 (7) Preparation of [Rh(TMPP)2][BF4] (3) from [Rh(TMPP)2CO][BF4] (7) .......................................... 107 (8) Reaction of [Rh(TMPP)2CO][BF4] (7) with small molecules .......................................................... 107 (i) N2, 02, C02, H2. ........................................... 107 (ii) CNR ................................................................ 107 (iii) pyridine. ......................................................... 108 B. X-ray Crystallography ........................................................... 108 (1) [Rh(TMPP)2(CO)2][BF4] oCHzClz (6). ..................... 109 (i) Data Collection and Reduction. .................... 109 (ii) Structure Solution and Refinement. ............ 109 (2) [Rh(nZ-TMPP)(TMPP)CO][BF4Jo2C6H5 (7). ............ .109 (i) Data Collection and Reduction. .................... 109 (ii) Structure Solution and Refinement. ............ 111 3. Results ............................................................................................... 111 A. Reactivity of [Rh(n3-TMPP)2][BF4]2 (3) with Carbon Monoxide. .............................................................................. 111 B. NMR Spectroscopic Studies. ................................................. 112 C. Electrochemistry of [Rh(nz-TMPP)(TMPP)CO][BF4] (7) and [Rh(TMPP)2(CO)2][BF4] (6). ...................................... 114 D. Crystal Structures of [Rh(TMPP)2(CO)n][BF4] (n=1,2) ...... 114 (1) [Rh(nZ-TMPP) (TMPP)CO][BF4] (7 ). ........................ 116 (2) [Rh(TMPP)2(CO)2][BF4] (6). .................................... 116 E. Solid State Reactions of (3) - (7 ) with Carbon Monoxide. .............................................................................. 120 4. Discussion .......................................................................................... 123 List of References ................................................................................... 131 CHAPTER V REVERSIBLE CARBON MON OXIDE ADDITION TO SOL-GEL DERIVED COMPOSITE FILMS CONTAINING THE MOLECULE [Rh(TMPP)2(CO)]1+ ......................................................... 133 1. Introduction ....................................................................................... 134 2. Experimental .................................................................................... 136 A. Preparation of Composite Films Containing [Rh(TMPP)2(CO)][BF4] (7) ................................................. 136 (1) Zirconia Films ............................................................ 136 (2) Titania Films .............................................................. 136 B. Flow Rate Determination ...................................................... 136 C. Electrochemical Measurements ............................................ 138 3. Results ............................................................................................... 138 4. Discussion .......................................................................................... 144 List of References ................................................................................... 145 CHAPTER VI CHEMISTRY OF [RhH(n3-TMPP)2][BF4]2 WITH ISOCYANIDE LIGANDS ................................................ 146 1. Introduction ....................................................................................... 147 2. Experimental .................................................................................... 147 A. Synthesis ................................................................................ 147 (1) Preparation of [Rh(TMPP)2(CNBut)2][BF4]2 (9) ................................................................................ 148 (2) Preparation of [Rh(TMPP)2(CNPri)2][BF4]2 (10) .............................................................................. 148 (3) Reaction of [Rh(n3—TMPP)2][BF4]2 (3) with other isocyanides ....................................................... 149 (i) Methyl isocyanide .......................................... 149 (a) one equivalent. .............................................. 149 (b) five equivalents. ............................................. 149 (ii) Cyclohexyl isocyanide. .................................. 150 (iii) n-Butyl isocyanide. ........................................ 150 (4) Reactions of [Rh(TMPP)2(CNBut)2][BF4]2 (9) ......... 150 (i) Cobaltocene: Preparation of [Rh(TMPP)2(CNBut)2][BF4] (ll) .................. 150 (ii) TMPP ............................................................. 151 (iii) tert-Butyl isocyanide. .................................... 151 (iv) Carbon monoxide. .......................................... 152 B. X-ray Crystallography ........................................................... 152 (1) [Rh(TMPP)2(CNBut)2][BPh4]2 (9) ............................ 152 (i) Data Collection and Reduction. .................... 152 (ii) Structure Solution and Refinement. ............ 154 3‘. Results ............................................................................................... 154 A. Synthesis and Spectroscopic Characterization of [Rh(TMPP)2(CNR)2][BF4]2 (R = But, Prl) ........................... 154 B. Magnetic and EPR Spectroscopic Properties of [Rh(TMPP)2(CNR)2][BF4]2 (R = But, Pri) ........................... 157 C. Redox Chemistry of [Rh(TMPP)2(CNR)2][BF4]2 (R = But, Pri) ................................................................................. 158 D. X-ray Crystal Structure of [Rh(TMPP)2(CNBut)2][BPh4] .............................................. 163 4. Discussion .......................................................................................... 166 List of References ................................................................................... 169 CHAPTER VII CHEMISTRY OF Rh(II) AND Rh(III) COMPLEXES LIGATED BY PHOSPHINO- PHEN OXIDE LIGANDS ................................................. 172 1. Introduction ....................................................................................... 173 2. Experimental .................................................................................... 173 A. Preparation of eq-[RhH(TMPP)(TMPP-O)][BF4] (12) .......... 174 (1) Reactions of [Rh(n3-TMPP)2][BF4]2 (3) with nucleophiles ............................................................... 17 4 (i) TMPP. ............................................................ 17 4 (ii) Tetra-n-butyl ammonium cyanide. ............... 17 4 (iii) Tricyclohexyl phosphine. .............................. 175 (2) Reduction ofax-[Rhmm3-TMPPXTI3-TMPP- 0)][BF4]2 (5) ............................................................... 175 B. Reaction of [Rh(n3-TMPP)2][BF4]2 (3) with excess TMPP .................................................................................... 175 C. Oxidation of eq-[RhH(TMPP)(TMPP-0)][BF4] (12); Synthesis of eq-[RhIH(n3-TMPP)(113-TMPP-0)][BF4J2 (l3) ......................................................................................... 176 D. Reaction of eq-[RhH(TMPP)(TMPP-0)][BF4] (12) with CO .......................................................................................... 176 (1) Infrared study. ........................................................... 176 (2) NMR study. ................................................................ 177 E. Dealkylation ofax-[Rh111(n3-TMPP)(n3-TMPP-O)] [BF4]2 (5): Formation of ax,eq- and ax,ax-[Rh(n3- TMPP-0)2][BF4] (l4) and (15) ............................................. 17 7 (1) TMPP .......................................................................... 177 (i) Bulk reaction. ................................................ 17 7 (ii) NMR reaction. ............................................... 178 (2) Tetra-n-butyl ammonium iodide ................................ 17 8 F. Reactions of ax-[Rhm(n3—TMPP)(n3-TMPP-0)] [BF412 (5) .............................................................................. 178 (1) HBF4oEt20 ................................................................ 178 (2) CO ............................................................................... 179 (3) H2 ................................................................................ 179 (4) MeOH ......................................................................... 179 3. Results ............................................................................................... 180 A. Synthesis and Characterization of eq- [RhH(TMPP)(TMPP-0)][BF4] (12) .................................... 180 B. Redox behavior of eq-[RhH(TMPP)(TMPP-0)][BF4] (l2) ......................................................................................... 182 C. Reactivity of ax-[Rhmm3-TMPP)(Ti3-TMPP-0)][BF4J2 (5) .......................................................................................... 185 D. Chemistry of eq-[RhH(TMPP)(TMPP-0)][BF4] (12) with CO ................................................................................. 189 4. Discussion .......................................................................................... 192 List of References ................................................................................... 197 CHAPTER VIII CHEMISTRY OF TRIS(2,4,6- TRIMETHOXYPHENYL)PHOSPHINE WITH RHODIUM (I) AND IRIDIUM (I) OLEFIN COMPLEXES ................................................................... 198 xiii 1. Introduction 199 2. Experimental 199 A. Synthesis 199 (1) Preparation of [Rh(cod)(n2-TMPP)][BF4] (16) .......... 200 (2) Preparation of Rh(cod)(n2-TMPP-O) (l7) .................. 201 (3) Reaction of [Rh(C2H4)2Cl]2 with TMPP ................... 201 (4) Preparation of Ir(cod)(n2-TMPP-0) (18) ................... 202 (5) Preparation of [Ir(TMPP)2(CO)2][BF4] (l9) ............. 202 (6) Reaction of TMPP with [Ir(TMPP)2(CO)2][BF4] (19) 203 (7) Oxidation of [Ir(TMPP)2(CO)2][BF4] (19) ................. 204 (i) Chemical oxidation with NOBF4 .................. 204 (ii) Chemical Oxidation with [(p- BX'CGH4)3N][BF4]. 204 (iii) Bulk Electrolysis. 205 (8) Reaction of [Ir(TMPP)2(CO)2][BF4] (19) with Iodine 205 B. X-ray Crystallography 207 (1) [Ir(TMPP)2(CO)2][BF4] (19) - CH2012 ..................... 207 (i) Data Collection and Reduction. .................... 207 (ii) Structure Solution and Refinement. ............ 209 3. Results 209 A. Synthesis and Characterization of [M(cod)(n2- TMPP)]1+ (M = Rh) and M(cod)(n2-TMPP-0) (M = Rh, Ir) 209 B. Synthesis and Spectroscopic Characterization of [Ir(TMPP)2(CO)2][BF4] (19) 213 C. Crystal Structure of [Ir(TMPP)2(CO)2][BF4] (19) ............... 215 D. Oxidation of [Ir(TMPP)2(CO)2][BF4] ................................... 215 4. Dionnum'nn 220 List of References 223 CHAPTER IX CONCLUSION 225 List of References 232 APPENDIX A PHYSICAL MEASUREMENTS 233 APPENDIX B CHEMISTRY OF TRIS(2,4,6- TRIMETHOXYPHENYL)PHOSPHIN E WITH 1_ Intrnrlnr‘finn 2. Experimental GROUP VI METALS 235 23.5 236 A. Synthesis 236 (1) Preparation of (n3-TMPP)MO(CO)3 (20) ................ (2) Reaction of TMPP with (n6-C7H3)W(CO)3 ............ (3) Reaction of TMPP with MoCl3(THF)3 ................... (4) Reaction of TMPP with M02Cl4(MeCN)4 .............. (i) 4 equivalents. (ii) 6 equivalents. B. X-ray Crystallography (1) (n3-TMPP)M0(CO)3 (20) . CH2012 ........................ (i) Data Collection and Reduction. ................. (ii) Structure Solution and Refinement. ......... 3. Results A. Synthesis and Characterization of (n3-TMPP)Mo(CO)3 (20) B. Molecular Structure of (n3-TMPP)M0(CO)3 (20) .............. C. Reaction of TMPP with (116-C7H8)W(CO)3 ....................... D. Reaction of TMPP with Solvated Molybdenum Chloride Complexes 4_ Di cmrqeinn List of References APPENDIX C COMPILATION OF 31P NMR SPECTRAL DATA ..... 236 237 237 237 238 238 238 238 240 240 240 244 247 247 248 25 1 ...252 10. 11. 12. LIST OF TABLES Page pKa values for methoxy substituted triphenylphosphines .................. 14 Summary of Crystallographic Data for TMPP (1) and TMPP=O - 2H20 (2) ................................................................................................. 37 Selected bond distances (A) and angles (deg) for TMPP (l). ............... 47 Dihedral angles (deg) between the planes of the aryl rings and the plane described by Cipso for TMPP (l) and TMPP=O . 2H20 (2). ........................................................................................................... 48 Selected bond distances (A) and angles (deg) for TMPP=O . 2H20 (2). ................................................................................................ 52 Summary of crystallographic data for [RhH(n3-TMPP)2][BF4]2 (3) and [RhHI(n3-TMPP)2][BF4][PF6]2 (4a). ........................................ 64 EPR Spectroscopic and Magnetic Susceptibility Data for [Rh(n3- TMPP)2][BF4]2 (3) ................................................................................. 74 Selected bond distances (A) and angles (deg) for [RhH(n3- TMPP)2][BF4]2 (3). ................................................................................ 79 Selected bond distances (A) and angles (deg) for [Rhm(n3- TMPP)2][PF6]2[BF4] . ........................................................................... 86 Summary of crystallographic data for [Rh(TMPP)2(CO)2][BF4] . CH2012 (6) and [Rh(TMPP)2(CO)]lBF4] ° ZCGHG (7). .......................... 110 Selected bond distances (A) and angles (deg) for [Rh(TMPP)2(CO)][BF4] . ZCGHG (7). .................................................... 119 Selected bond distances (A) and angles (deg) for [Rh(TMPP)2(CO)2][BF4] . CH2C12 (6). ................................................ 122 13. 14. 15. 16. 17. 18. 19. Summary of crystallographic data for [RhH(TMPP)2(CNBut)2] [BPh4]2 (9a). .......................................................................................... 153 Selected bond distances (A) and angles (deg) for [RhH(TMPP)2(CNBut)2][BPh4]2 (9a). ................................................... 165 Summary of crystallographic data for [IrI(TMPP)2(CO)2][BF4] . CH2012 (19) ............................................................................................ 208 Selected bond distances (A) and angles (deg) for [IrI(TMPP)2(CO)2][BF4] - CH2C12 (l9). ............................................... 217 Summary of crystallographic data for (n3-TMPP)Mo(CO)3 . CH2C12 (20). ........................................................................................... 239 Selected bond distances (A) and angles (deg) for (n3-TMPP)- M0(CO)3 (20). ......................................................................................... 246 Compilation of 31P NMR spectral data ................................................ 252 xvii 10. 11. 12. 13. LIST OF FIGURES Page Schematic drawing of tris(2,4,6-trimethoxyphenyl)phosphine ........... 12 Tolman cone angles for various tertiary phosphines. .......................... 15 111, n2 and n3 bonding modes of TMPP. ................................................ 17 Plot of the cone angle versus the v(CO)A1 stretch for various Ni(CO)3(PR3) complexes. ...................................................................... 43 ORTEP representations of the two crystallographically independent molecules of TMPP. .......................................................... 46 ORTEP diagram of TMPP=O . 2H20. .................................................. 50 Packing diagram of TMPP=O . 2H20. ................................................. 51 EPR spectrum of [Rh(n3-TMPP)2][BF4]2 (3) in a CHgClz/Me- THF glass at 77 K. ................................................................................. 72 ORTEP representation of the [Rh(n3-TMPP)2]2+ (3) molecular cation with 40% probability ellipsoids. Phenyl ring atoms are Shown as small spheres of arbitrary size for clarity. ........................... 76 ORTEP drawing emphasizing the coordination sphere of [Rh(n3- TMPP)2]2+. ............................................................................................. 77 Cyclic voltammogram of [RhII(n3-TMPP)2][BF4]2 (3) in 0.1 M TBABF4 in CH2C12. ............................................................................... 81 Cyclic voltammogram of [RhHI(n3-TMPP)2][BF4]3 (4) in 0.1 M TBABF4 in Cchlz. ............................................................................... 83 ORTEP drawing of the [RhIH(n3-TMPP)2]3+ (4) molecular cation. Phenyl ring atoms are shown as small spheres of arbitrary size for clarity. ............................................................................................... 85 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 1H and 31P NMR spectra of ax-[Rhm(TMPP)(TMPP-0)][BF4]2 in CD2012. 31P NMR spectrum of the product from the reduction of [Rhn(n3- TMPP)2][BF4]2 (3) with cobaltocene Cyclic voltammograms in 0.1 M TBABF4/CH2C12 for (a) [RhIII(n3-TMPP)2][BF4]3 (4), (b) [Rh(TMPP)2(CO)2][BF4] (6), (c) [Rh(TMPP)2(CO)][BF4] (7). ORTEP representation of [Rh(TMPP)(n2-TMPP)(CO)]1+ (7 ). ............. PLUTO drawing of [Rh(TMPP)(n2-TMPP)(CO)]1+ (7) emphasizing the coordination geometry about the Rh atom. .............. ORTEP drawing of [Rh(TMPP)2(CO)2]1+ (6) with 30% probability ellipsoids. Proposed pathway for the reversible reaction between [Rh11(n3- TMPP)2]2+ (3) and CO. High yield synthetic routes to compounds 3-7. .................................... Schematic diagram of the apparatus used to estimate CO concentrations. Infrared spectra of a zirconia composite film of [Rh(TMPP)2(CO)][BF4] (7): (a) in the absence of CO. (b) exposed to CO. Electronic absorption spectra of (a) [Rh(TMPP)2(CO)][BF4] and (b) [Rh(TMPP)2(CO)2][BF4] in: (I) CH2C12. (II) a zirconia composite film. ' Cyclic voltammograms of a zirconia composite film containing [Rh(TMPP)2(CO)][BF4] (7) and LiCF3SO3 cast onto a platinum disk electrode: (a) under a N2 atmosphere. (b) after purging the cell for 30 seconds with CO. (c)-(e) after flushing the CO saturated cell with N2 for 2, 5 and 10 min. EPR spectrum of [RhH(TMPP)2(CNBut)2][BF4]2 (9) in the solid- state at 100 K. EPR spectrum of [RhH(TMPP)2(CNBut)2][BF4]2 (9) in a CHZClZ/Me-THF glass at 100 K. 88 93 115 117 118 121 124 125 137 139 141 143 159 160 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. EPR spectrum of [RhH(TMPP)2(CNPri)2][BF4]2 (10) in a CH2C12/Me-THF glass at 100 K. ........................................................... 161 ORTEP representation of the molecular cation [Rh(TMPP)2(CNBut)2]2+ in structure (9). Thermal ellipsoids are Shown at the 40% probability level. Carbon and oxygen atoms of the phosphine ligand are shown as small spheres of arbitrary Size for clarity ......................................................................................... 164 EPR Spectrum of eq-[RhII(TMPP)(TMPP-0)][BF4] (12) at 100 K in a 1:1 CH2C12/Me-THF glass .............................................................. 181 Cyclic voltammograms of ax-[RhHI(n3-TMPP)(n3-TMPP- 0)][BF412 (5) and eq-[RhH(TMPP)(TMPP-O)][BF4] (12) in 0.1 M TBABF4/CH2012. ................................................................................... 183 Possible dealkylation products resulting from nucleophilic attack on ax-[Rhm(n3-TMPP)(T]3-TMPP-0)][BF4]2 (5). ................................. 187 EPR spectrum of a frozen 1:1 CH2C12/Me-THF solution of eq- [RhII(TMPP)(TMPP-O)][BF4] (12) exposed to CO. .............................. 191 Reaction scheme depicting the relationship between the various isomers of dealkylated and non-dealkylated complexes 3, 4, 5, 12 and 13. .................................................................................................... 194 Schematic drawing of the electrolysis cell used in the electrochemical oxidation of [IrI(TMPP)2(CO)2]1+ (l9). ...................... 206 ORTEP representation of the molecular cation [IrI(TMPP)2(CO)2]1+ (19). ...................................................................... 216 Synthetic pathways to Rh-TMPP complexes ....................................... 227 n1 - n3 bonding modes for TMPP and [TMPP-01 ................................ 228 Variable temperature 1H NMR spectra of (n3-TMPP)M0(CO)3 (20) in d8-toluene. .................................................................................. 242 Variable temperature 1H NMR Spectra of (n3-TMPP)M0(CO)3 (20) in CD2C12. ....................................................................................... 243 ORTEP representation of (n3-TMPP)Mo(CO)3 (20) ............................. 245 42. Proposed pathway for intramolecular exchange in (113- TMPP)M0(CO)3 (20). 249 Ag/AgCl br But ca. cm-1 cod esd EtOH tram LIST OF SYMBOLS AND ABBREVIATIONS Angstrtim Silver-silver chloride reference electrode aryl group broad tert-butyl circa, about centimeter wavenumber 1,5-cyclooctadiene cyclic voltammetry cyclohexyl degree centigrade doublet (NMR), day, deuterated parts per million (ppm) bis(dicyclohexylphosphino)ethane doublet of doublets anodic peak potential cathodic peak potential electron paramagnetic resonance electromagnetic unit estimated standard deviation ethyl ethanol molar extinction coefficient Fast Atom Bombardment epr g-value, gram Gauss hour xxii TMPP-H (M)Hz NBA NMR ox PCy3 PPh3 P(mes)3 Pp_m Prl psi red r.t. s sh SQUID tris(2,4,6-trimethoxyphenyl)phosphonium (Mega)Hertz infrared Kelvin wavelength medium moles per liter methyl acetonitrile methanol milligram minute milliliter millimole multiplet bridging ligand Bohr magneton microliter nanometer frequency 3-nitrobenzyl alcohol nuclear magnetic resonance oxidation tricyclohexylphosphine triphenylphosphine trimesitylphosphine parts per million isopropyl pounds per square inch reduction room temperature singlet (NMR), strong (IR) shoulder Superconducting Quantum Interference Device triplet TBABF4 THF TMPP TMPP=O TMPP-0 TMPP-CH3 TMPP-CH2CI tetra-n-butylammonium tetrafluoroborate tetrahydrofuran tris(2,4,6-trimethoxyphenyl)phosphine triS(2,4,6—trimethoxyphenyl)phosphine oxide P[{C6H2(CH3O)3}2{C6H2(CH3O>2O}]' tris(2,4,6-trimethoxyphenyl)methyl phosphonium tris(2,4,6-trimethoxyphenyl)chloromethyl phosphonium tetramethylsilane ultraviolet Volt versus, very strong weak halide ligand xxiv LIST OF COMPOUNDS (l) TMPP (2) TMPP=O (’3) [Rh11(n3-TMPP)2][BF4]2 (4) [RhHI(n3-TMPP)2][BF4]3 (5) ax-[Rh111(n3-TMPP)(113-TMPP—O)][BF4]2 (6) [RhI(TMPP)2(CO)2][BF4] (7) [RhI(TMPP)2(CO)l[BF4] (8) [RhI(TMPP)2(CO)(CNBut)] [BF 4] (9) [Rh11(TMPP)2(CNBut)2][BF4]2 (10) [RhH(TMPP)2(CNPri)2][BF412 (11> [RhI(TMPP)2(CNBut)2][BF4] (12) eq-[RhH(TMPP)(TMPP-0)][BF4] (13) eq-[RhHI(n3-TMPP)(113-TMPP-0)][BF4]2 (14) ax,eq-[Rhm(n3-TMPP-0)2][BF4] (15) ax,ax-[Rhm(n3-TMPP-0)2][BF4] (16) [Rh1(cod)(n2-TMPP)][BF4] (17) Rh1(cod)(n2-TMPP-0) (18) IrI(cod)(’n2-TMPP-O) (l9) [IrI(TMPP)2(CO)2][BF4] (20) (n3-TMPP)M00(CO)3 CHAPTER I INTRODUCTION 2 A. Role of Tertiary Phosphines in Transition Metal Chemistry The use of tertiary phosphines as ancillary ligands has played a prominent role in the development of modern coordination and organo- metallic chemistry. This statement is most evident from the observation that phosphine complexes are known for virtually all transition metals.1 These versatile ligands are capable of stabilizing a variety of metals in a range of oxidation states. Furthermore, transition metal-phosphine complexes, particularly those of the later transition metals, catalyze a number of industrially important organic processes.2 These range from olefin hydrogenation, hydroformylation, hydrosilation and hydrocyanation to polymerization and oligomerization of olefins and acetylenes. In addition to catalytic processes, transition metal phosphine complexes are also capable of performing a number of stoichiometric organic transformations.3 An important aspect in the use of transition metal-phosphine complexes in catalysis has been the potential for controlling catalyst activity and selectivity by modifying the substituent groups of the phosphine. By varying the R groups of a phosphine, both the electronic and steric properties of the ligand may be changed. This in turn will affect the reactivity properties of the metal to which the phosphine is bound. It is important to realize, however, that steric and electronic effects are not independent of each another; often, substitution of one of the R substituents results not only in a change in the steric properties of ligands, but also the electronic properties as well. For instance as the size of a phosphine is increased, the C-P-C angle is forced to expand which affects the percentage of S-character in the phosphorus lone pair and hence the phosphine donor properties. Furthermore, sterically hindering substituents also affect the ability of the phosphine to interact strongly with the metal center and therefore influence __——_—'—___4 3 the metal phosphorus bond strength. As a consequence of this interrelationship, there is considerable interest in quantifing and separating the factors that influence transition metal-phosphorus bonding. The electronic properties of a particular phosphine molecule are dominated by the donor capacity of the phosphorus lone pair of electrons Which is quite sensitive to the nature of the phosphine substituents, or R groups. By altering the R groups, and therefore the donor ability of the phosphine, the electron density at the metal center may be tuned in order to increase or decrease the reactivity of the complex. The electronic donating properties of phosphine ligands are typically divided into C and 1: contributions.4 Generally, the 6 contribution dominates and 1: contributions are only evident for phosphines with electronegative substituents such as -F, -OR, -Cl and -OAr. The n-accepting behavior of phosphines has been debated for many years. Conventional wisdom has traditionally maintained that the n-acceptor capability is facilitated through back donation from filled metal d-orbitals into empty low-lying phosphorus- 3d orbitals,5 but, recent theoretical calculations suggest that the frontier orbitals responsible for n-interactions consist primarily of phosphorus 3p character having local 6* symmetry with no involvement of phosphorus-3d orbitals.6 Highly electronegative substituents, such as -F, Cl, and OR, result in a highly polar P-X bond, which in turn, lowers the energy of the 6* orbitals, making them more accessible for metal Tr-donation. Unfortunately, it is difficult to separate 6 and 1: contribution due to their inherent synergistic relationship.7 Nonetheless, the importance of understanding metal-phosphine bonding has prompted the search for methods to quantify the individual contributions. Early methods of measuring phosphine donor strength were performed by measuring the 4 energy of the A1 carbonyl stretching vibration for the monosubstituted Ni carbonyl complex, Ni(CO)3PR3.8 The energy of this band provides an indication of the total donor ability (0' + 1:) of the phosphine. The more strongly donating a phosphine, the more electron density resides at the metal center. Consequently, the degree of n-accepting of the carbonyl groups will increase, resulting in a lowering of the carbonyl stretching frequencies. Therefore, strongly donating phosphines will result in lower values of v(CO) for Ni(CO)3L. Another commonly used measure of phosphine donor strength is pKa.9 Values of pKa for the conjugate acid R3PH+ provide an indication of the o donor strength of the phosphine. However, it is often argued that these values are not a good measure of ligand donor strength because pKa is a measure of the phosphine's affinity for a hard acid and not a soft transition metal. Furthermore, variations in phosphine size may result in differing solvation energies that will effect the pKa. In spite of these considerations, pKa values have been found to correlate well with other experimental observations regarding phosphine basicity.1°aib Others have sought more quantitative methods to measure ligand 6 and 1: effects; these have met with mixed success.11 In particular, Giering et al. have divided phosphine ligands into three classes based upon the correlation of oxidation potential for the phosphine complex 11 5- MeCpMn(CO)2PR3 and phosphine pKa values:113 1 s I (cs-donor / n—donor) PR3 (R = Et, Bu, Cy, etc.) 1 II (cs-donor only) 5 PR3 (R = Me), PPh3_n(R)n (n = 1, 2; R = Me, Et, Bu) P(p-X-Ph)3 (X = H, Me, OMe) Class III (cs-donor / n-acceptor) P(OR)3 R = Et, Ph, Pri, Me P(p-X-Ph)3 (X = F, Cl) Similar correlations were found between pKa, on°, vco, cone angle and heats of reactions for other systemslla'f Although Giering classified phosphine ligands in terms of rt-acidity and n-basicity, it is important to stress that these n-effects are relatively small compared to those observed for true 1:- acids and bases such as CO or Cl'. Furthermore, the classification system is specific to the particular system under investigation, although in practice, these class boundries differ only slightly. In addition to varying the electronic properties of the phosphine, substitution of the R groups often affects the size of the ligand. Altering the size can have major ramifications on the subsequent chemistry of the phosphine, as in the case of phosphines with unusually large steric requirements, such as tricyclohexylphosphine, tri-tert-butylphosphine, tri-o- tolylphosphine and trimesitylphosphine; these ligands are capable of kinetically stabilizing metal complexes with low coordination numbers, thereby engendering electronically unsaturated metal centers that are highly reactive.12 For example, bulky phosphines have been used to isolate a series of two-coordinate, 14-electron complexes of Pt0 and Pd0.13 Although these systems are formally electron deficient, their reactivity is governed by the degree of steric overcrowding that is present.14 6 Another example of the use of bulky phosphines to stabilize electronically unsaturated metal centers is provided by the recent work of Kubas and co-workers who used sterically hindering phosphines to create coordinatively unsaturated Mo0 and W0 complexes with the general formula trans-M(CO)3(PR3)2 (R = cyclohexyl, iso-propyl).15 These 5-coordinate complexes are stabilized by an "agostic" interaction of a C-H bond on one of the phosphines with the metal center. This interaction is readily displaced in favor of small substrates including N2 and H2. One consequence of steric crowding is that these complexes tend to undergo intramolecular C-H bond activation to form metallated phosphine complexes.16,17 Whitesides and co-workers have used steric effects to generate a highly reactive "Pt0L2" fragment in situ by reductive elimination of neopentane from cis-PtII(dCype)H(CH2C(CH3)3) (dCype = bis(dicyclohexylphosphino)ethane, (Cy)2P(CH2)2P(Cy)2) (eq 1)18 14-electron complex activates a variety of aryl Cy.P.Cy">1’ A cy‘ ,Cy \ R pt ————> Pt° + CH3 (1) and aliphatic C-H bonds, but does not undergo intramolecular C-H activation, due, presumably, to the bent configuration of the dCype moiety, which precludes interaction of the C-H bonds of the cyclohexyl groups and affords the substrate better access to the metal center.18 The steric interactions of phosphines is also implicated in the activity of many homogenous catalysts by facilitating ligand dissociation that opens up coordination Sites for an incoming substrate. Mechanistic studies of Wilkinson's catalyst, Rh(PPh3)Cl, have shown that initial dissociation of 7 phosphine is a key step in olefin hydrogenation.19 The size of the ancillary ligands also imparts selectivity to a number of reactions, a factor that is particularly important in the development of homogeneous catalysts for asymmetric synthesis. Although it is certainly true that both steric and electronic properties of phosphine play a role in the reactivity and catalytic behavior of transition metal phosphine complexes, steric effects tend to dominate. In order to address the size of a phosphine ligand, the concept of cone angle, (9, was developed by Tolman.20 This method, which is based on rigid space filling models, defines the cone angle of a phosphine as the apex angle of a solid cone, centered 2.28 A from the phosphorus atom, that encompasses the phosphine substituents atoms at their van der waal radii. Cone angles thus described have been found to correlate well with a number of spectroscopic and physical properties of transition metal phosphine complexes.21 A more realistic description of ligand size was provided by the development of cone angle profiles.22 This method, based on X-ray crystallographic data, accounted for the clefts and intermeshing of the phosphine substituents and gave a better indication of the true ligand requirements. B. Development of Ether-Phosphines An important aspect in any homogenous catalytic process is the availability of accessible coordination sites in which an incoming substrate molecule may bind.23 Typical catalysts employing tertiary phosphine complexes rely on the steric bulk of the phosphine to promote ligand dissociation in order to create a coordinately unsaturated metal center. Another strategy is to incorporate weak donor atoms (e.g. solvent molecules) into the metal coordination sphere that can be easily displaced in favor of the incoming substrate. For example, recent work with zirconium 8 cyclopentadienyl systems has shown that replacement of a chloride ion with THF in the coordination sphere creates a highly reactive cationic metal center, capable of performing a variety of organic transformations.24 A number of compounds that incorporate solvent into their coordination sphere have been found to be quite effective hydrogenation, hydroformylation and polymerization catalysts.25 Furthermore, although many catalyst precursors use olefins as supporting ligands, under catalytic conditions, the olefins are hydrogenated and replaced by solvent to form the active catalyst (eq 2).26 1 R3P\ /\—_| + solv R3P\ ,SOlV—l +1 M p + 2 H2 —> ’M\ (2) 12,3]?I \/ R3P SOIV An alternative approach to the design of a new catalyst with open coordination sites is to structurally modify the ancillary phosphine ligands by incorporating weak donor groups directly into the phosphine. A variety of functional groups that are capable of acting as weak donors have been combined with phosphines to form polydentate ligands;27 these provide additional stabilization to the complex by chelating to the metal center, but because of their weak donor nature these groups are easily displaced in favor of the incoming substrate. After the transformations are complete and the product dissociates, the tethered donor atom is able to quickly reassociate.28 As a result of this "arm on/arm off' behavior, these functionalized ligands have been described as being hemi-labile.29 A further advantage of a tethered donor group, particularly an oxygen donor atom, is the observed rate enhancement for oxidative addition reactions, which are well recognized as important steps in many catalytic processes.30 These reactions are facilitated 9 by the presence of the oxygen atom, which provides anchimeric assistance to the oxidative addition of polar substrates. The successful implementation of phosphine ligands bearing oxygen donor substituents to catalysis has been demostrated in a number of important reactions.27 For example, Shells Higher Olefin Process (SHOP) uses a nickel complex ligated by the phosphino-carboxylate ligand, thPCH2C02H, to selectively polymerize ethylene in the presence of other olefins.31 Further work in this area was carried out by Knowles and co-* workers who developed a series of chiral phosphines (shown below) that incorporate ortho substituted phenyl groups.32 Cationic rhodium complexes supported by these chiral ligands successfully hydrogenate pro-chiral olefins, an important step in the synthesis of cr-amino acids, achieving enantiomeric excesses > 95%. The asymmetric hydrogenation of eMePQ cm; «at; o no MeO pamp camp dipamp . (N-acylamino)-cinnamic acid catalyzed by the diphosphine complex [Rh(cod)(diPAMP)]1+ (diPAMP = (R,R)-1,2-bis[2-methoxyphenyl)phenyl- phosphino]ethane), a key step in Monsanto's L-Dopa synthesis, represents the first commercialized catalytic asymmetric process.33 The development of these catalysts ignited a revolution in the area of asymmetric hydrogenation promoted by transition metal complexes. Although the role of the ortho- methoxy group is primarily steric in nature, replacement of the methoxy substituent with sterically equivalent groups resulted in lower catalytic 10 activity, suggesting that the methoxy groups provide an electronic influence as well.34 Although a variety of donor groups, ranging from amines and thioethers to enolates and diketonates, have been used as functional groups on phosphine ligands, ether donors are the most prevalent.27a Being uncharged, ethers are very weak donors, which renders them highly highly labile, particularly for the soft late transition metals, with which they are relatively incompatible. The first crystallographic evidence for the coordinating ability of a pendent ether group was reported for the Rh(III) arsine complex RhCl3[MeAs(o-CGH4(OMe)]2.35 Later in the mid 1970's, research on the chemistry of ether-phosphines expanded due to the independent work of Roundhill, Rauchfuss and Shaw.36 These groups were able to show that ortho-substituted phenylphosphines were capable of binding a variety of late transition metals through both the oxygen and phosphorus atoms. Rauchfuss found that the coordinated methoxy groups were loosely bound and could dissociate readily, thus generating vacant coordination sites. As a result, many of these complexes exhibit stereochemically non-rigid behavior and reversible substrate addition. More recently, Lindner and co-workers have developed a comprehensive series of ether-phosphines in which the ether group is tethered by an ethylene group as opposed to a rigid phenyl ring.37 In addition to being the first to use a more flexible linkage in a ether-phosphine ligand, Lindner has also pioneered the use of chiral and cyclic ether donors such as THF and dioxane. His work has focused on the use of these chelating phosphines for the development of catalysts that assist in the carbonylation of methanol to acetic acid. These phosphino-ether complexes, like others in 11 the same category, exhibit fluxional behavior, reversible substrate binding, and enhanced oxidative addition and reductive elimination chemistry promoted by the on/off nature of the ether donors. Others have found similar results with both mono- and di- phosphines containing pendent ether substituents.33 C. Chemical Aspects of Tris(2,4,6-trimethoxyphenyl)phosphine Although the use and development of ether-phosphines is extensive, largely due to the recent efforts of Lindner et al., none of the aforementioned systems have addressed the combined effects of high basicity and steric bulk on complex stability and reactivity. We are interested in developing the coordination chemistry of ether-phosphine ligands that combine steric bulk and strong donor capability with chelating ability to stabilize complexes in highly reactive and uncommon oxidation states. To this end, we have undertaken the comprehensive study of the coordination chemistry of the highly basic and sterically hindering ether-phosphine, tris(2,4,6- trimethoxyphenyl)phosphine (Figure 1). The molecule tris(2,4,6-trimethoxyphenyl)phosphine, which we refer to as TMPP, was originally prepared by Soviet chemists in the late 1950's.39 The phosphine later reappeared in literature in the mid 1980's, when Wada and co-workers described its extraordinarily high basicity (PKa = 11.2) and steric properties (cone angle ~ 184°).40 Wada has exploited the phosphine's unusual basic properties in a variety of organic transformations, including mild ring opening reactions of terminal epoxides and facile dealkylations.“=1 More recently, the basicity and solubility properties of its phosphonium salts have been applied in the extraction of metal ions.42 The unusual properties of this phosphine are derived from the presence of the methoxy groups in the ortho and para positions. The electron releasing nature of the methoxy 12 I Me M Figure 1. Schematic drawing of tris (2,4,6-t1imethoxyphenyl) phosphine (TMPP). — 13 substituents increases the nucleophilicity of the phosphorous lone pair via the mesomeric effect of the phenyl rings. The affect of sequential methoxy substitution at the ortho and para ring positions on the phosphine's basicity is Shown in Table 1.421) As the number of methoxy substituents is increased, there is an increase in the basicity. Substitution of a -OMe group in the para position has a greater effect on the basicity than substitution in the ortho positions. In addition, the presence of two or more methoxy groups on one ring results in a dramatic increase in nucleophilicity of the phosphine. For example, the presence of one 2,6-dimethoxy substituted ring results in a higher basicity (pKa = 5.39) than either of the tris 2-ortho- or para- substituted phosphines (pKa = 4.47 and 4.75). In any event, there is an overall increase in phosphine basicity as the number of substituted aryl groups are increased. The electron donating effect of the methoxy groups is dramatically illustrated by the comparison of the pKa of TMPP with that of the unsubstituted phosphine, triphenylphosphine (pKa = 2.73).43 In addition to increasing the nucleophilicity, the presence of the methoxy groups in the 2 and 6 positions also serves to augment the steric bulk of the ligand. The cone angle of TMPP was reported by Wada and later confirmed by our group to be approximately 184°. Comparision with other phosphines reveals that TMPP is one of the most sterically encumbering phosphines known (Figure 2). The cone angle of TMPP is nearly 40° larger than that of triphenylphosphine, commonly considered a sterically hindering ligand (cone angle = 145°).20 As evidenced by the results of numerous studies, sterically encumbering ligands lend kinetic stability to normally very reactive transition metal centers, thus creating electron deficient complexes. Such complexes would be expected to exhibit unusual reactivity not normally observed for complexes ligated by more conventional ligands. 14 Table l. PKa values for methoxy substituted triphenylphosphines (from reference 42b). 2-MeO 4-MeO 2,6-MeO 2,4,6—Me0 (x-MeOPh)3P 4.47 4.75 9.33 11.2 (x-MeOPh)2PPh 4.01 4.06 7.28 8.22 (x-MeOPh)PPh2 3.33 3.67 5.39 5.77 Cone Angle - P(mesityl)3 200° 4 180° — - PButa 160° - - PBu3s 140° - 120°- Figure 2. Tolman cone angles for various tertiary phosphines. 16 Perhaps most importantly for the purpose of our research goals, the presence of ether donors in the ortho positions provide the opportunity for multidentate coordination modes for the phosphine. Unlike other ether- phosphines that contain only one ether substituents, the presence of methoxy groups in the ortho positions of all three aryl rings allow for TMPP to participate in 111 through 113 bonding arrangements (Figure 3). Most other ether-phosphines typically have only one ether donor and therefore may only participate in n1 or 112 type bonding. The multiple chelating ability of TMPP will be of paramount importance in this work for the stabilization of coordinatively unsaturated, electronic deficient, metal diphosphine complexes. Moreover, the multidentate capabilities afford versatility, by facilitating adjustments to the electronic requirements dictated by the metal center. As the electronic requirements of the metal change, TMPP compensates by altering its bonding mode. As a result, the phosphine may accommodate a number of different metal oxidation states. The physical and structural properties of TMPP and its corresponding oxide are fully described in Chapter 2. The combination of the chelate effect provided by the multidentate capability of TMPP, together with the kinetic stability afforded by the steric Size of the ligand, render TMPP an excellent ligand to stabilize highly reactive metal centers in unusual oxidation states. Moreover, unlike other transition metal complexes stabilized by bulky phosphines, the presence of hemi-labile ether donors will provide these complexes with open coordination sites in which further chemistry, particularly with small substrates, may take place. Interestingly, with the exception of our work, only four reports of transition metal TMPP complexes have appeared in the literature.“ Figure 3. 17 I Me 0 Me. ’0 0’ Me \ ”Ar' Ari O,Me M? P 0' Me Ni—PWAr' Me’O OYMe 0‘Me n3 111, n2 and n3 bonding modes of TMPP. 18 D. Chemistry of Mononuclear Rh(II) Complexes Based on these considerations, our goal is to exploit the unique properties of TMPP in order to stabilize, isolate and study the chemistry of odd electron transition-metal systems. We are particularly interested in the chemistry of mononuclear d7 rhodium complexes. A survey of the literature reveals that the bulk of rhodium coordination and organometallic chemistry is dominated by the catalytically important +1 and +3 oxidation states.45 These oxidation states form the basis for oxidative addition and reductive elimination reactions that are key steps in many important catalytic processes.46 In contrast to the ubiquity of mononuclear Rh(I) and Rh(III) complexes, the chemistry of the paramagnetic, divalent oxidation state remains relatively unexplored. Generally, these species exist only as highly reactive, fleeting intermediates or as impurities in the chemistry of d6 and d8 rhodium.47 The lack of stability of mononuclear Rh(II) complexes is a result of the proclivity of these systems to (a) dimerize to form a metal-metal bonded species and (b) disproportionate to the more stable Rh(I) and Rh(III) oxidation states. These factors notwithstanding, a number of the Rh(II) metallo-radicals are known.43 Under suitable conditions, sterically encumbering phosphines react with Rh(III) trihalides in alcohol to yield paramagnetic Rh(II) compounds formulated as trans-RhClz(PR3)2 (PR3: PCy3, P(o-tolyl)3, P(But)2Me).49 With a few exceptions, these complexes are poorly characterized, prone to decomposition and invariably contaminated with diamagnetic hydrido species. As a result, there has been no comprehensive investigation of the chemistry of these paramagnetic species undertaken and only a few studies have appeared.50 In fact, only recently, the first structural studies of such 19 species were finally reported.51,52 Paramagnetic mononuclear species of rhodium that are coordinated by less sterically demanding ligands are also known; these complexes , however, are ligated by "non-innocent" ligands in which extensive delocalization of the unpaired electron is possible.56 Typically, such species are not considered to be authentic metallo-radicals. In addition to the aforementioned phosphine complexes, a variety of meta-stable mononuclear Rh(II) complexes have been generated by electrochemical oxidation and reduction of Rh(III) and Rh(I) species.53 Few stable mononuclear Rh(II) organometallic complexes have been reported; many of these exist as short-lived intermediates and hence are poorly characterized.54 A recent exception is the unprecedented Rh(II) dialkyl Species Rh(2,4,6-Pri3CeH2)2(tht)2 (tht = tetrahydrothiophene) prepared by Wilkinson and co-workers from Rthl3(tht)3.55 This complex is one of the rare examples of a paramagnetic Rh(II) complex to be structurally characterized by X-ray crystallography. . By far the most promising results for Rh(II) complexes have been found in the realm of Rh(II) porphyrin chemistry. Highly reactive, short-lived Rh(II) Species have been implicated in a number of unusual and fundamental organometallic reactions involving mononuclear and dinuclear rhodium porphyrin complexes.57 Wayland and co-workers have reported strategies that favor the existence of these paramagnetic intermediates that involve introducing bulky substituents onto the porphyrins that discourage metal- metal bond formation through steric interactions. By studying such key systems, they were able to isolate Rh(II) radicals that reversibly couple carbon monoxide to form dimetal-a-diketones (eq 3 - 5).58 Through the use of even more sterically demanding porphyrins, dimerization of the carbonyl radicals is disfavored and a 17 electron Rh(II) monocarbonyl species may be 20 (por)Rhm-CH3 h" r (por)RhH- (3) benzene 0 || (por)RhII - + CO ‘ ‘ (por)Rhn’ C ° (4) O 0 II II 2 (pOI‘)RhII’ C 0 V — (130191111111, C\ C ’ Rhm(p or) (5) ll (por) = tetramesitylporphyrin 0 observed directly.“c In contrast to other 17 electron d7 carbonyl complexes, the nonlinear RhCO’ unit behaves as an acyl radical, undergoing reactions at the carbon and not at the metal center. Even more fascinating are the recent reports that these Rh(II) porphyrin systems reversibly and selectively activate methane to form Rh(III)-hydride and Rh(III)-alkyl Species (eq 6).59 Thermodynamic data point to the presence of a linear four-centered transition state that precludes the activation of aromatic C-H bonds due to steric inhibition by the porphyrin ligands. As a result, these Rh(II) porphyrin systems are highly selective towards the activation of alkyl versus aryl C-H bonds. Recent results by Wayland and co-workers indicate that these Rh(II) metalloradicals also react with acrylates to produce C-C bonded oligomers.60 CH4 III (por)RhIII -C H3 2 (por)RhH° :2“— (P0r>Rh--,C\--H--Rh(por)= + (6) If H (P0P)Rhm -H (por) = tetramesitylporphyrin or tetraxylylporphyrin In light of these recent results, the study of rhodium based odd electron systems is becoming an important area of research. In general, these Species 21 must be afforded kinetic stabilization by sterically encumbering ligands that preclude dimerization. In the case of Rh(II), further stabilization must be provided by electronic factors that favor the +2 oxidation state relative to the +1 and +3 oxidation states, hence avoiding undesirable disproportionation reactions. The work in this thesis supports the use of TMPP as a good candidate for providing the proper combination of kinetic and thermodynamic stability for the isolation of mononuclear Rh(II) complexes. As a backdrop for this work we note that Shaw and co—workers had earlier demonstrated that mixed phosphorus and oxygen donor chelate ligands are capable of stabilizing d7 metal centers.61 Reactions of P(But)3(o-MeOC6H4) with MCl3 oxHZO (M = Rh, Ir) in refluxing alcohol produced the paramagnetic mononuclear Rh(II) and Ir(II) complexes, M(P(But)2(o-OC6H4))2 (eq 7). During the reaction the metals are reduced from M(III) to M(II) and the phosphine undergoes dealkylation to form a phosphino-phenoxide chelating ligand. The Ir complex represents one of the few crystallographically characterized examples of a stable Ir(II) mononuclear species. The remarkable stability of these species illustrates the potential for other bulky mixed P,O chelates to stabilize odd 0\ 132 P i0H M0130 xHZO + 2 R2P—Q ’—> 0 M’ I) (7) A I \ P o Moo 2 M = Rh, Ir R = Me, tBu electron systems. This thesis reports the results of investigations concerning the isolation and reactivity of paramagnetic Rh(II) centers using the triaryl phosphine ligand TMPP. Our entry into the area of mononuclear d7 rhodium 22 chemistry began with the synthesis of the bis-phosphine complex [cis-Rh(n3- TMPP)2][BF4]2 from a reaction between TMPP and the solvated dinuclear salt [Rh2(MeCN)10][BF4]4. The details of the synthesis and characterization of this metalloradical is presented in chapter 3 together with its associated redox chemistry. Chapter 4 examines in detail the reversible CO chemistry that was observed for [cis-Rh(n3—TMPP)2][BF4]2. This rather complicated chemistry involves the disproportionation of [RhH(n3-TMPP)2]2+ to give Rh(I) carbonyl and Rh(III) species that upon loss of CO, recombine to form the original Rh(II) species. The reversible CO behavior of the Rh(I) carbonyl species is examined more closely in chapter 5, wherein we describe the potential application of this process for the development of a CO sensing composite material by incorporation the Rh(I) carbonyl species into a porous sol-gel derived glass. As an extension of the reversible CO chemistry discussed in chapter 3, reactions of [Rh(n3-TMPP)2][BF4]2 with isocyanide ligands is presented in chapter 6. Unlike the complex electron transfer chemistry that was observed with CO, the weaker n-acidity of isocyanides allows for the isolation of paramagnetic Rh(II) diisocyanide complexes, rare examples of stable organometallic Rh(II) complexes. The redox properties and reactivity of these organometallic species are also discussed. Chapter 7 examines the effect of nucleophiles on the stability of Rh(II) and Rh(III) bis- TMPP complexes. Nucleophilic attack on [Rh(n3-TMPP)2][BF4]2 results in dealkylation of a coordinated methoxy substituent to yield a Rh(II) complex stabilized by a phosphino-phenoxide interaction. The chemistry of this species is discussed along with its relationship to analogous Rh(III) compounds formed from dealkylation reactions of [Rh(n3-TMPP)2][BF4]3. Chapter 8 discusses the formation of Rh(I) and Ir(I) olefin species with TMPP and the subsequent formation of the Ir(I) dicarbonyl cation 23 [Ir(TMPP)2(CO)2]1+, which, unlike its rhodium counterpart, is stable with respect to CO loss. Attempts to oxidize this complex to yield an Ir(II) species are also presented. Finally, the application of the n3-bonding mode to group VI metal tricarbonyl fragments is presented in Appendix B. Specifically, the isolation and characterization of the highly fluxional molecular species (113- TMPP)Mo(CO)3 is discussed. 10. 11. 24 List of References McAuliffe, C. A. In Comprehensive Coordination Chemistry; Wilkinson, G.; Gillard, R. D.; McCleverty, J. A., eds.; Pergamon: Oxford, 1987, chapter 14, p 989. (a) Pignolet, L. H. Homogeneous Catalysts with Metal Phosphine Complexes; Plenum: New York, 1983. (b) Parshall, G. W. Homogeneous Catalysis: The Applications and Chemistry of Catalysis by Soluble Transition Metal Complexes; Wiley: New York, 1980. (c) Comils, B. In New Syntheses with Carbon Monoxide; Falbe, J ., ed.; Springer-Verlag: Berlin, 1980. (d) Catalytic Aspects of Metal Phosphine Complexes, Advances in Chemistry Series 196; Alyea, E. C.; Meek, D. W. Eds.; American Chemical Society: Washington, D. C. 1982. (e) Hughes, R. P. in Comprehensive Organometallic Chemistry; Wilkinson, G.; Stone, F. G. A.; Abel, E. W., eds.; Pergamon: New York, 1982; vol. 5, p 228. Collman, J. P.; Hegedus, L. 8.; Norton, J. R.; Finke, R. G. Principles and Applications of Organotransition Metal Chemistry; University Science Books: Mill Valley, Ca. 1987. Cotton, F. A.; Wilkinson, G.; Advanced Inorganic Chemistry, 5th ed.; Wiley: New York, 1988, pp 64-68. For instance, see Huheey, J. E. Inorganic Chemistry; Harper and Row: New York, 1983; p 832. (a) Marynick, D. S. J. Am. Chem. Soc. 1984, 106, 4064. (b) Xiao, S.; Trogler, W. 0.; Ellis, D. E.; Berkovitch-Yellin, Z. J. Am. Chem. Soc. 1983, 105, 7033. (c) Whango, M. H.; Steward, K. R. Inorg. Chem. 1982, 21,1720. Honeychuck, R. V.; Hersh, W. H. Inorg. Chem. 1987,26, 1826. Tolman, C. A. J. Am. Chem. Soc. 1970,92, 2953. (a) Allman, T.; Goel, R. G. Can. J. Chem. 1982,60, 716. (b) Henderson, W. A.; Streuli, C. A. J. Am. Chem. Soc. 1960,82, 5791. (c) Streuli, C. A. Anal. Chem. 1960,32, 985. (d) Schuster-Woldan, H. G.; Basolo, F. J. Am. Chem. Soc. 1966, 88, 1657. (a) Golovin, M. N.; Rahman, Md. M.; Belmonte, J. E.; Giering, W. P. Organometallics 1985,4, 1981. (b) Rahman, Md. M.; Liu, H.-Y.; Prock, A.; Giering, W. P. Organometallics 1987, 6, 650. (c) Dahlinger, K.; Falcone, F.; P6e, A. J. Inorg. Chem. 1986, 25, 2654. (a) Rahman, Md. M.; Liu, H.-Y.; Prock, A.; Giering, W. P. Organometallics 1989,8, 1. (b) Eriks, K.; Giering, W. P.; Liu, H.-Y.; Prock, A. Inorg. Chem. 1989,28, 1759. (c) Liu, H.-Y.; Eriks, K.; Prock, 12. 13. 14. 15. 16. 17. 25 A.; Giering, W. P. Organometallics, 1990,9, 1758. (d) P6e, A. J. Pure Appl. Chem. 1988, 60, 1209. (e) Zizelman, P. M.; Amatore, C.; Kochi, J. K. J. Am. Chem. Soc. 1984,106, 3771. (f) Nolan, S. P.; Hoff, C. D. J. Organomet. Chem. 1985, 290, 365. (g) Bush, R. C.; Angelici, R. J. Inorg. Chem. 1988, 27, 681. (h) Bodner, G. M.; May, M. P.; McKinney, L. E. Inorg. Chem. 1980, 19, 1951. (i) Colton, R.; Tomkins, I. B., Aust. J. Chem. 1967, 19, 1143. (j) Graham, W. A. G. Inorg. Chem. 1968, 7, 315. (k) Treichel, P. M. Inorg. Chem. 1968, 7, 1942. Clark, H. C.; Hampden-Smith, M. J. Coord. Chem. Rev. 1987, 79, 229. (a) Immirzi, A.; Musco, A. A. J. Chem. Soc., Chem. Commun. 1974, 400. (b) Immirzi, A.; Musco, A.; Zambelli, P.; Carturan, G. Inorg. Chem. Acta 1975, 13, L13. (c) Matsumoto, M.; Yoshida, H.; Nakatsu, N.; Yoshida, T.; Otsuka, S. J. Am. Chem. Soc. 1974, 96, 3322. (d) Moynihan, K. J .; Chieh, C.; Goel, R. G. Acta Cryst. 1979, B35, 3060. (e) Otsuka, S.; Yoshida, T.; Matsumoto, M.; Nakatsu, K. J. Am. Chem. Soc. 1976,98, 5850. (f) Yoshida, T.; Yamagata, T.; Tulip, T. H.; Ibers, J. A.; Otsuka, S. J. Am. Chem. Soc. 1978,100, 2063. (a) Clark, H. C.; Hampden-Smith, M. J. J. Am. Chem. Soc. 1986, 108, 3829. (b) Clark, H. C.; Ferguson, G.; Hampden-Smith, M. J .; Kaitner, B.; Ruegger, H. Polyhedron, 1988, 7, 1349. (c) Forries, J .; Green, M. L. H.; Spencer, J. L.; Stone, F. G. A. J. Chem. Soc, Dalton Trans. 1977 , 1006. (d) Yoshida, T.; Otsuka, S. J. Am. Chem. Soc. 1977, 99, 2134. (a) Kubas, G. J. J. Chem. Soc, Chem. Commun. 1980, 61. (b) Kubas, G. J .; Ryan, R. R.; Swanson, B. 1.; Vergamini, P. J .; Wasserman, H. J. J. Am. Chem. Soc. 1984, 106, 451. (c) Kubas, G. J.; Ryan, R. R.; Wrobleski, D. A.J. Am. Chem. Soc. 1986, 108, 1339. (d) Wasserman, H. J.; Kubas, G. J.; Ryan, R. R. J. Am. Chem. Soc. 1986, 108, 2294. (e) Kubas, G. J .; Unkefer, C. J .; Swanson, B. I.; Fukushima, E. J. Am. Chem. Soc. 1986, 108, 7000. (f) Kubas, G. J .Acc. Chem. Res. 1988, 21, 120. (g) Kubas, G. J. Comments Inorg. Chem. 1988, 7, 17. Examples of cyclometallation in Pt0 and Pd0 complexes: (3) Clark, H. C.; Goel, A. B.; Goel, R. G.; Ogini, W. J. Organomet. Chem. 1978, 157, 219. (b) Clark, H. C.; Goel, A. B.; Goel, S. Inorg. Chem. 1979,18, 2803. For reviews of cyclometallation reactions see: (c) Omae, I. Coord. Chem. Rev. 1980,32, 235. (d) Dehand, J .; Pfeffer, M. Coord. Chem. Rev. 1976, 18, 327. (e) Bruce, M. I. Angew Chem, Int. Ed. Engl. 1977, 16, 73. (f) Parshall, G. W. Acc. Chem. Res. l970,3, 139. (a) Crabtree, R. H.; Holt, E. M.; Lavin, M.; Morehouse, S. M. Inorg. Chem. 1985, 24, 1986. (b) Shaw, B. L. J. Am. Chem. Soc. 1975,97, 3856. (c) Jones, C. E.; Shaw, B. L.; Turtle, B. L. J. Chem. Soc, Dalton Trans. 1974, 992. (d) Cheney, A. J .; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1972, 754. 18. 19. 20. 21. 22. 23. 24. 25. 26 (a) Hackett, M.; Whitesides, G. M. J. Am. Chem. Soc. 1988, 110, 1449. (b) Hackett, M.; Ibers, J. A.; Whitesides, G. M. J. Am. Chem. Soc. 1988, 110, 1436. (c) Hackett, M.; Ibers, J. A.; Jernakoff, P.; Whitesides, G. M. J. Am. Chem. Soc. 1986, 108, 8094. (Ca) Halpern, J. Inorg. Chem. Acta 1981,,50 11. (b) Halpern, J; Wong, S. J. Chem. Soc, Chem. Commun. 1973, 629. (c) Halpern, J.; Okamoto, T.; Zakhariev, A. J. Mol. Catal. 1977, 2, 65. (a) Tolman, C. A. Chem. Rev. 1977, 77, 313. McAuliffe, C. A. In Comprehensive Coordination Chemistry; Wilkinson, G.; Gillard, R. D.; McCleverty, J. A., eds.; Pergamon: Oxford, 1987, chapter 14, p 1027. (a) Ferguson, G.; Alyea, E. C.; Dias, S. A.; Restivo, R. J. Inorg. Chem., 1977, 16, 2329. (b) Ferguson, G.; Alyea, E. C.; Khan, M. Inorg. Chem. 1978, 17, 2965. (c) Ferguson, G.; Alyea, E. C., Dias, S. A.; Pavez, M. Inorg. Chem. Acta. 1979,37, 45. (d) Aylea, E. C.; Ferguson, G.; Somogyvari, A. Inorg. Chem. 1982, 21 , 1369. (e) Aylea, E. C.; Ferguson, G.; Siew, P. Y. Can. J. Chem. 1983, 61, 257. (f) Ferguson, G.; Siew, P. Y.; Goel, A. B. J. Chem. Res. 1979, 4337. (g) Richardson, J. F.; Payne, N. C. Can. J. Chem. 1977,55, 3203. (h) Smith, J. D.; Oliver, J. D. Inorg. Chem. 1978, 17, 2585. Crabtree, R. H. Comments Inorg. Chem. 1985, 4, 229. (a) Guram, A. S.; Jordan, R. F.; Taylor, D. F. J. Am. Chem. Soc. 1991, 113, 1833. (b) Jordan, R. F.; Bradley, P. K.; Baenziger, N. C.; LaPointe, R. E. J. Am. Chem. Soc. 1990, 112, 1289. (c) Jordan, R. F.; LaPointe, R. E.; Baenziger, N.; Hinch, G. D. Organometallics, 1990,9, 1539. (d) Jordan, R. F.; Taylor, D. F.; Baenziger, N. C. Organometallics 1990, 9, 1546. (e) Guram, A. S.; Jordan, R. F. Organometallics 1990, 9, 2190. (f) Jordan, R. F.; Guram, A. S. Organometallics 1990, 9, 2116. (g) Jordan, R. F.; Bradley, P. K.; LaPointe, R. E.; Taylor, D. F. New J. Chem. 1990, 14, 505. (h) Jordan, R. F.; LaPointe, R. E.; Bradley, P. K.; Baenziger, N. Organometallics 1989, 8, 2892. (i) Jordan, R. F.; Taylor, D. F. J. Am. Chem. Soc. 1989, 111, 778. 0) Jordan, R. F.; Bajgur, C. S.; Dasher, W. E.; Rheingold, A. L. Organometallics 1987, 6, 1041. (k) Jordan, R. F.; Bajgur, C. S.; Willet, R.; Scott, B. J. Am. Chem. Soc. 1986,108, 7410. (1) Jordan, R. F.; LaPointe, R. E.; Bajgur, C. S.; Echols, S. F.; Willett, R. J. Am. Chem. Soc. 1987, 109, 4111. (a) Sen, A. Acc. Chem. Res. 1988, 21, 421 and references therein. (b) Sen, A.; Lai, T.-W. Organometallics, 1984,3, 866. (c) Sen, A.; Lai, T.- W. Organometallics, 1982, 1, 415. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 27 (a) Crabtree, R. Acc. Chem. Res. 1979, 12, 331 and references therein. (b) Schrock, R. R.; Osborn, J. A. J. Am. Chem. Soc. 1976, 98, 2134. (c) Schrock, R. R.; Osborn, J. A. J. Am. Chem. Soc. 1976,98, 2143. For recent reviews on functionalized phosphines see: (a) Bader, A.; Lindner, E. Coord. Chem. Rev. 1991, 108, 27. and references therein. (b) Rauchfuss, T. B. Homogeneous Catalysis with Metal Phosphine Complexes; L. Pignolet, Ed.: Plenum, New York, 1983, 239. (a) Homer, L.; Simons, G. Z. Naturforsch. 1984, 39b, 504. (b) Homer, L.; Simons, G. Z. Naturforsch. 1984, 39b, 512. Jeffrey, J. C.; Rauchfuss, T. B. Inorg. Chem. 1979,18, 2658. (a) Miller, E. M.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1974, 480. (b) Empsall, H. D.; Hyde, E. M.; Jones, C. E.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1974, 1980. (a) Keim, W.; Behr, A.; Gruber, B.; Hoffmann, B.; Kowaldt, F. H.; Kiirschner, U.; Limbacker, B.; Sistig, F. P. Organometallics 1986, 5, 2356 and references therein. (b) Keim, W.; Kowaldt, F. H.; Goddard, R.; Kruger, C. Angew. Chem., Int. Ed. Engl. 1978, 17, 466. (c) Freitas, E. R.; Gum, C. R. Chem. Engin. Progr. 1979, 75, 73. (d) Peuckart, M.; Keim. W. Organometallics 1983,2, 594. (a) Knowles, W. S. Acc. Chem. Res. 1983,16, 106 and references therein. (b) Knowles, W. S.; Sabacky, M. J .; Vineyard, B. D. Adv. Chem Ser. 1974, 132, 274. (c) Knowles, W. S.; Sabacky, M. J .; Vineyard, B. D. J. Chem. Soc., Chem. Commun. 1972, 10. (a) Vineyard, B. D.; Knowles, W. S., Sabacky, M. J .; Bachman, G. C.; Weinkauff, D. J. J. Am. Chem. Soc. 1977, 99, 5946. (b) Knowles, W. S.; Sabacky, M. J .; Vineyard, B. D.; Weinkauff, D. J. J. Am. Chem. Soc. 1975, 97, 2567. Studies relavent to the mechanism of asymmetric hydrogenation by Rh-dipamp complexes: (a) Landis, C. R.; Halpern, J. J. Am. Chem. Soc. 1987, 109, 1746. (b) Halpern, J. Pure & Appl. Chem. 1983,55, 99. (c) Brown, J. M.; Maddox, P. J. J. Chem. Soc., Chem. Commun. 1987, 1276. (d) Alcock, N. W.; Brown, J. M.; Derome, A. E.; Lucy, A. R. J. Chem. Soc., Chem. Commun. 1985, 575. (e) Brown, J. M.; Cutting, I. J. Chem. Soc., Chem. Commun. 1985, 578. (D Brown, J. M.; Chaloner, P. A.; Morris, G. A. J. Chem. Soc., Chem. Commun. 1983, 664. (g) Brown, J. M.; Parker, D. J. Chem. Soc., Chem. Commun. 1980, 342. (h) Brown, J. M.; Chaloner, P. A. J. Chem. Soc., Chem. Commun. 1980, 344. Graziani, R.; Bombieri, G.; Volponi, L.; Panattoni, C. J. Chem. Soc. (A) 1969, 1236. 36. 37. 38. 28 (a) Mason, R.; Thomas, K. M.; Empasall, H. D.; Fletcher, S. R.; Heys, P. N.; Hyde, E. M. Jones, C. E; Shaw, B. L. J. Chem. Soc., Chem. Comm. 1974, 612. (b) Jones, C. E.; Shaw, B. L.; Turtle, B. L. J. Chem. Soc., Dalton Trans. 1974, 993. (c) Rauchfuss, T. B.; Patino, F. T.; Roundhill, D. M. Inorg. Chem. 1975, 14, 652. ((1) Jeffrey, J. C.; Rauchfuss, T. B. Inorg. Chem. 1979, 18, 2658. (a) Lindner, E.; RothfuB, H.; Fawzi, R.; Hiller, W. Chem. Ber. 1992, 125, 541. (b) Lindner, E.; Bader, A.; Mayer, H. A. Z. Anorg. Allg. Chem. "1991, 598/599, 235. (c) Lindner, E.; Dettinger, J .; Mockel, A. Z. Naturforsch. 1991, 46b, 1519. (d) Lindner, E.; Dettinger, J. Z. Naturforsch. 199], 46b, 432. (e) Lindner, E.; Reber, J.-P. Z. Naturforsch. 199], 46b, 1581. (f) Lindner, E.; Karle, B. Chem. Ber. 1990, 123, 1469. (g) Lindner, E.; Norz, H. Chem. Ber. 1990, 123, 459. (h) Lindner, E.; Bader, A.; Braunling, H.; Jira, R. J. Mol. Catal. 1990, 57, 291. (i) Lindner, E.; Fawzi, R.; Mayer, H. A.; Eichele, K.; Pohmer, K. J. Organomet. Chem. 1990, 386, 63. (j) McCann, G. M.; Carvill, A.; Lindner, E.; Karle, B.; Mayer, H. A. J. Chem. Soc., Dalton Trans. 1990, 3107. (k) Lindner, E.; Karle, B. Chem. Ber. 1990, 123, 1464. (l) Lindner, E.; Karle, B. Z. Naturforsch. 1990, 45b, 1108. (m) Lindner, E.; Speidel, R. Z. Naturforsch. 1989, 44b, 437. (n) Lindner, E.; Norz, H. Z. Naturforsch. 1989, 44b, 1493. (o) Lindner, E.; Reber, J. P.; Wegner, P. Z. Naturforsch. 1988,43b, 1268. (p) Lindner, E.; Andres, B. Z. Naturforsch. 1988, 43b, 369. (q) Lindner, E.; Schober, U.Inorg. Chem. 1988,27, 212. (r) Lindner, E.; Andres, B. Chem. Ber. 1988,12], 829. (s) Lindner, E. ; Sickinger, A.; Wegner, P. J. Organomet. Chem. 1988, 349, 75. (t) Lindner, E.; Meyer, S. J. Organomet. Chem. 1988, 339, 193. (u) Lindner, E.; Andres, B. Chem. Ber. 1987, 120, 761. (v) Lindner, E.; Schober, U.; Fawzi, R.; Hiller, W.; Englert, U.; Wegner, P. Chem. Ber. 1987, 120, 1621. (w) Lindner, E.; Meyer, 8.; Wegner, P.; Karle, B.; Sickinger, A.; Steger, B. J. Organomet. Chem. 1987, 335, 59. (x) Lindner, E.; Schober, U.; Stangle, M. J. Organomet. Chem. 1987,33], 013. (y) Lindner, E.; Schrober, U.; Glaser, E.; Norz, H.; Wegner, P. Z. Naturforsch. 1987, 42b, 1527. (z) Lindner, E.; Scheytt, C.; Wegner, P. J. Organomet. Chem. 1986,308, 311. (aa) Lindner, E.; Scheytt, C. Z. Naturforsch. 1986, 41b, 10. (bb) Lindner, E.; Mayer, H. A.; Wegner, P. Chem. Ber. 1986,119, 2616. (cc) Lindner, E.; Sickinger, A.; Wegner, P. J. Organomet. Chem. 1986,312, C37. (dd) Lindner, E.; Raudeler, H.; Scheytt, C.; Mayer, H. A.; Hiller, W.; Fawzi, R.; Wegner, P. Z. Naturforsch. 1984, 39b, 632. (a) Mason, M. R.; Verkade, J. G. Organometallics 1992, 11 , 1514. (b) , Mason, M. R.; Verkade, J. G.; J. Am. Chem. Soc. 1991,113, 6309. (0) Mason, M. R.; Su, Y.; Jacobson, R. A.; Verkade, J. G. Organometallics 1991, 10, 2335. (d) Werner, H.; Stark, A.; Shulz, M. ; Wolf, J. Organometallics 1992, 11 , 1126. (e) Werner, H.; Hampp, A.; Peters, K.; Peters, E. M.; Walz, L.; von Schnering, H. G. Z. Naturforsch. 1990, 45b, 1548. (D Anderson, G. K.; Kumar, R. Inorg. Chem. 1984, 23, 4064. (g) Anderson, G. K.; Corey, E. R.; Kumar, R. Inorg. Chem. 1987, 26, 97. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 29 (h) Anderson, G. K.; Kumar, R. Inorg. Chim. Acta. 1988, 146, 89. (i) Alcock, N. W.; Platt, A. W. G.; Pringle, P. G. J. Chem. Soc., Dalton Trans. 1989, 2069. (j) Reddy, V. V. S.; Whitten, J. E.; Redmill, K. A.; Varshney, A.; Gray, G. M. J. Organomet. Chem. 1989, 372, 207. (k) Reddy, V. V. S.; Varshney, A.; Gray, G. M. J. Organomet. Chem. 1990, 391, 259. (a) Protopopov, I. S.; Kraft, M. Y. Zhumal Obshchei Khimii 1963,33, 3050. (b) Protopopov, I. S.; Kraft, M. Y. Med. Prom. SSSR, 1959, 13, 5; Chem. Abstr. 1960, 54, 10914c. (a) Wada, M.; Higashizaki, S.; J. Chem. Soc., Chem. Commun. 1984, 482. (b) Wada, M.; Higashizaki, S.; Tsuboi, A. J. Chem. Res. 1985, (S), 38; (M), 0467. (a) Wada, M.; Tsuboi, A. J. Chem. Soc., Perkin Trans. I 1987, 151. (b) Wada, M.; Tsuboi, A.; Nishimura, K.; Erabi, T. Nippon Kagaku Kaishi 1987, 7, 1284. (a) Yamashoji, Y.; Matsushita, T.; Wada, M.; Shono, T. Chem. Lett. 1988, 43. (b) Yamashoji, Y.; Matsushita, T.; Tanaka, M.; Shono, T.; Wada, M. Polyhedron 1989, 8, 1053. Streuli, C. A. Anal. Chem. 1960, 32, 985. (a) Bowmaker, G. A.; Cotton, J. D.; Healy, P. C.; Kildea, J. D.; Silong, S. B.; Skelton, B. W.; White, A. H. Inorg. Chem. 1989,28, 1462. (b) Li, G. Q.; McAuliffe, C. A.; Mackie, A. G.; Mac Rory, P. P.; Ndifon, P. T. J. Chem. Soc., Dalton Trans. 1992, 1297. (c) Kurosawa, H.; Tsuboi, A.; Kawasaki, Y.; Wada, M. Bull. Chem. Soc. Jpn. 1987, 60, 3563. (d) Baker, L.-J.; Bowmaker, G. A.; Camp, D.; Effendy,; Healy, P. C.; Schmidbauer, H.; Steigelmann, 0.; White, A. H.Inorg. Chem. 1992, 31, 3656. (a) Hughes, R. P. in Comprehensive Organometallic Chemistry; Wilkinson, G.; Stone, F. G. A.; Abel, E. W., eds.; Pergamon: New York, 1982; vol. 5. Collman, J. P.; Hegedus, L. 8.; Norton, J. R.; Finke, R. G. Principles and Applications of Organotransition Metal Chemistry; University Science Books: Mill Valley, Ca. 1987; chapter 5 and references therein. Busetto, C.; D‘Alfonso, A.; Maspero, F.; Perego, G.; Zazzetta, A. J. Chem. Soc., Dalton Trans. 1977, 1733. For a review of Rh(II) complexes see Felthouse, T. R. Prog.Inorg. Chem. 1982,29, 73 and references therein. 49. 50. 51. 52. 53. 54. 30 (a) Bennett, M. A.; Lonstaff, P. A. J. Am. Chem. Soc., 1969,91, 6266. (b) Moers, F. G.; DeJong, J. A. M.; Beaumont, P. M. J. J. Inorg. Nucl. Chem. 1973,35, 1915. (c) Masters, C.; Shaw, B. L. J. Chem. Soc. (A) 1971, 3679. (d) Empsall, H. D.; Heys, P. N.; Shaw, B. L. Transition Met. Chem. 1978, 3, 165. (e) Empsall, H. D.; Hyde, E. M.; Pawson, D.; Shaw, B. L.; J. Chem. Soc., Dalton Trans. 1977 , 1292. (a) Vleck, A. Inorg. Chim. Acta 1980, 43, 35. (b) Holah, D. G.; Hughes, A. N.; Hui, B. C. Can. J. Chem. 1975, 53, 3669. (c) Valentim', G.; Braca, G.; Sbrana, G.; Colligiani, A. Inorg. Chim. Acta 1983, 69, 215. (d) Valentini, G.; Braca, G.; Sbrana, G.; Colligiani, A. Inorg. Chim. Acta 1983, 69, 221. (a) Harlow, R. L.; Thorn, D. L.; Baker, R. T.; Jones, N. L. Inorg. Chem. 1992, 31 , 993. (b) Rappert, T.; Wolf, J .; Schulz, M.; Werner, H. Chem. Ber. 1992,125, 839. The crystal structure of RhIIClz(PPh3)2 was reported recently, Ogle, C. A.; Masterman, T. C.; Hubbard, J. L. J. Chem. Soc., Chem. Commun. 1990, 1733, but it was later shown to be that of the Rh(I) complex Rh(PPh3)2(CO)Cl, Dunbar, K. R.; Haefner, S. C. Inorg. Chem. 1992, 31 , 0000. Examples of coordination and organometallic Rh(II) complexes prepared electrochemically: (a) Cooper, S. R.; Rawle, S. C. ; Yagbasan, R.; Watkin, D. J. J. Am. Chem. Soc. 1991,113, 1600. (b) Rawle, S. C.; Yagbasan, R.; Prout, K.; Cooper, S. R. J. Am. Chem..Soc. 1987, 109, 6181. (c) Blake, A. J .; Gould, R. D. ; Holder, A. J .; Hyde, T. I.; Schroder, M. J. Chem. Soc., Dalton Trans. 1988, 1861. ((1) Anderson, J. E.; Gregory, T. P. Inorg. Chem. 1989, 28, 3905. (e) Scwarz, H. A.; Creutz, C. Inorg. Chem. 1983,22, 207. (e) Mulazzani, Q. G.; Emmi, S.; Hoffman, M. Z.; Venturi, M. J. Am. Chem. Soc. 1981, 103, 3362. (f) Kalle, U.; Klaui, W. Z. Naturforsch. 199], 46b, 75. (g) Bianchini, C.; Laschi, F.; Ottaviani, M. F.; Peruzzini, M.; Zanello, P.; Zanobini, F. Organometallics 1989,8, 893. (h) Bianchini, C.; Meli, A.; Laschi, F.; Vizza, F.; Zanello, P. Inorg. Chem. 1989,28, 227. (i) Pilloni, G.; Schiavon, G.; Zotti, G.; Zecchin, S. J. Organomet. Chem. 1977, 134, 305. (i) Fischer, E. 0.; Lindner, H. H. J. Organomet. Chem. 1964, 1, 307. (k) Fischer, E. 0.; Wawersik, H. J. Organomet. Chem. 1966,5, 559. (1) Keller, H. J .; Wawersik, H. J. Organomet. Chem. 1967,8, 185. (m) Dessy, R. E.; King, R. B.; Waldrop, M. J. Am. Chem. Soc. 1966,88, 5112. (n) Dessy, R. E.; Kornmann, R.; Smith, C.; Hayter, R. J. Am.‘ Chem. Soc. 1968,90, 2001. (a) Bianchini, C.; Laschi, F.; Ottaviani, F.; Peruzzini, M.; Zanello, P. Organometallics 1988, 7, 1660. (b) Bianchini, C.; Laschi, F.; Meli, A.; Peruzzini, M.; Zanello, P.; Frediani, P. Organometallics 1988, 7, 2575. 55. 56. 57. 58. 59. 60. 61. 31 Hay-Motherwell, R. S.; Koschmieder, S. U.; Wilkinson, G.; Hussain- Bates, B.; Hursthouse, M. B. J. Chem. Soc., Dalton Trans. 1991, 2821. (a) Billig, E.; Shupack, S.I.; Waters, J.H.; Williams, R.; Gray HE. J. Am. Chem. Soc. 1964, 86, 926.(b) Pneumatikakis, G.; Psaroulis, P. Inorg. Chim. Acta 1980, 46, 97. (c) Pandey, K. K.; Nehete, D. T.; Sharma, R. B. Polyhedron 1990, 9, 2013. (d) Peng, S. M.; Peters, K.; Peters, E. M.; Simon, A. Inorg. Chim. Acta 1985, 101, L35. (a) Wayland, B. B.; Woods, B. A.; Coffin, V. L. Organometallics 1986, 5, 1059. (b) Del Rossi, K. J .; Wayland, B. B. J. Am. Chem. Soc. 1985, 107, 7941. (c) Paonessa, R. S.; Thomas, N. C.; Halpern, J. J. Am. Chem. Soc. 1985,107, 4333. (d) Kadish, K. M.; Yao, C.-L.; Anderson, J. E.; Cocolios, P. Inorg. Chem. 1985,24, 4515. (e) Anderson, J. E.; Yao C.-L.; Kadish, K. M. Organometallics 1987, 6, 706. (f) Anderson, J. E.; Yao, C.-L.; Kadish, K. M. J. Am. Chem. Soc. 1987, 109, 1106. (g) Aoyama, Y.; Yoshida, T.; Sakurai, K.—I.; Ogoshi, H. Organometallics 1986, 5, 168. (a) Wayland, B. B.; Sherry, A. E.; Coffin, V. L. J. Chem. Soc., Chem. Commun. 1989, 662. (b) Wayland, B. B.; Sherry, A. E. J. Am. Chem. Soc. 1989, 111, 5010. (c) Wayland, B. B.; Sherry, A. E.; Poszmik, G.; Bunn, A. G. J. Am. Chem. Soc. 1992,114, 1673. (a) Wayland, B. B.; Ba, S.; Sherry, A. E. J. Am. Chem. Soc. 199], 113, 5305. (b) Sherry, A. E.; Wayland, B. B. J. Am. Chem. Soc. 1990,112, 1259. (a) Bunn, A. G.; Wayland, B. B. J. Am. Chem. Soc. 1992, 114, 6917. (b) Poszmik, G; Wayland, B. B. presented at the 203rd National meeting of the American Chemical Society, San Francisco, CA, April 5-10, 1992; paper INOR 67. (a) Empsall, H. D.; Hyde, E. M.; Jones, C. E.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1974, 1980. (b) Mason, R.; Thomas, K. M.; Empasall, H. D.; Fletcher, S. R.; Heys, P. N.; Hyde, E. M. Jones, C. E; Shaw, B. L. J. Chem. Soc., Chem. Comm. 1974, 612. (c) Empsall, H. D.; Hyde, E. M.; Shaw, R. L. J. Chem. Soc., Dalton Trans. 1975, 1690. (d) Empsall, H. D.; Heys, P. N .; McDonald, W. 8.; Norton, M. 0.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1978, 1119. CHAPTER II STRUCTURAL AND SPECTROSCOPIC PROPERTIES OF TRIS(2,4,6- TRIMETHOXYPHENYL)PHOSPHINE AND TRIS(2,4,6- TRIMETHOXYPHENYL)PHOSPHINE OXIDE 32 33 1. Introduction Before delving directly into the coordination chemistry of tris(2,4,6- trimethoxyphenyl)phosphine (TMPP), a summary of its spectroscopic and structural properties is in order. The initial synthetic report describing TMPP and several related quaternary phosphonium salts did not report any spectroscopic information about the phosphine; in fact only melting points and elemental analyses were given.1 Later, Wada described a more conventional route for the preparation of TMPP along with an account of its 1H NMR and infrared spectra] data.2 In addition to providing some spectral data for TMPP and several of its quaternary phosphonium salts, Wada estimated the pKa of the phosphine to be 11.2 by titration of the protic phosphonium salt with a series of bases and monitoring the process by 1H NMR spectroscopy. Later the value was redetermined by non-aqueous titrimetry in nitromethane and found to be 1102.3 The high nucleophilicity of TMPP is readily apparent from its observed chemical reactivity. Although some of the physical and spectroscopic properties of TMPP appeared during the course of our studies, a complete structural and spectroscopic study of this ligand has not yet been published to our knowledge.4 In this chapter, we describe the results of such studies completed in our laboratories. In addition to the spectroscopic and structural properties of TMPP, we are also interested in those of the corresponding phosphine oxide, in particular the affect that phosphine oxidation has on the spectroscopic and structural properties. Transition metal catalyzed oxidations of phosphines are not uncommon, particularly in the chemistry of late transition metal- phosphine complexes.5 An example of such a process was investigated in our own laboratories. In this work, TMPP was reacted with FeCl3 in the presence of oxygen to form a TMPP oxide-FeCl3 complex.6 In order to identify the I “I.“ 34 possible formation of any phosphine oxide by-products, it is necessary to study their physical properties, thus the oxide derivative of TMPP was prepared and fully characterized by spectroscopic and crystallographic methods. 2. Experimental A. Synthesis ( 1) Preparation of Tris(2,4,6-trimethoxyphenyl)phosphine (TMPP) TMPP was prepared by using a modification of the method previously described by Wada.2b A 1 liter 3-necked flask equipped with a condenser, a gas inlet, and a 120 mL addition funnel, was charged with 1,3,5- trimethoxybenzene (42.05 g, 0.25 mol) and diethyl ether (150 mL). The resulting solution was cooled to 0°C with an ice bath and a 2.5 M solution of butyllithium in hexanes (100 mL) was slowly added dropwise over a period of 1-2 h. The resulting suspension was stirred at 0°C for a total of 12 h. After this time, triphenylphosphite (19.65 mL, 0.075 mol) was slowly added dropwise to the suspension at 0°C. While stirring for 10-12 h, the reaction was allowed to slowly warm to r.t. The solvent was decanted off and the remaining white solid was filtered in air and washed with a minimal amount of cold EtOH (2 x 10 mL). The product was recrystallized in air by dissolving the solid in a large volume of hot EtOH (1 L), followed by filtration of the hot solution and concentration on a rotary evaporator. The resulting white crystalline solid was filtered, washed with cold EtOH (2 x 10 mL) and diethyl ether (4 x 10 mL), and dried in vacuo; yield, 26-28 g (65-70% based on P(OPh)3). IR (Nujol, CsI) cm'1: 1596, 1581, 1398, 1321, 1293, 1223, 1202, 1182, 1154, 1120, 1086, 1036, 949, 919, 816, 805, 789, 675, 634, 474. Electronic absorption spectrum 35 (CHgCN) kmax, nm (e): 276 (21.6 x 103), 245 sh (~29.3 x 103), 214 (87.2 x 103). 1H NMR (CD30N) 5 ppm: 3.45 (s, 18H, o-OCH3), 3.75 (s, 9H, p-OCH3), 6.07 (d, 4JP-H = 2.51 Hz, 6H, m-H). 31P (0D30N) 5 ppm: -66.3 (s). 13C (CD30N) 5 ppm: 109.19 (d, 1JP-C = 22.78 Hz, 01), 163.01 (d, 02,6), 92.21 (3, 03,5), 164.12 (8, C4), 56.42 (s, o-OCH3), 55.80 (s, p-OCH3). Cyclic Voltammetry (0.1 M TBABF4/ CH3CN, vs Ag / Ag01): EM = + 0.49 v, Ep,c(chemical) = — 0.09 V. (2) Preparation of Tris(2,4,6-trimethoxyphenyl)phosphine Oxide (TMPP=O) An excess of a 30% aqueous solution of H202 (1 mL) was added to a solution of TMPP (1.00 g, 1.98 mmol) in 10 mL of acetone. The colorless solution was gently refluxed for 8 h, after which time the solution was cooled and the volume was reduced to approximately 5 mL under reduced pressure. Slow addition of diethyl ether (50 mL) effected the precipitation of a large amount of white crystalline solid. The product was collected by suction filtration, washed with diethyl ether (4 x 5 mL) and dried in vacuo for 1 h; yield, 0.897 g (87%). IR (Nujol, CsI), cm'1: 1596, 1579, 1406, 1333, 1293, 1230, 1206, 1184, 1159, 1120, 1096, 1028, 952, 921, 812, 789, 692, 665, 640, 594, 527, 520, 501, 482, 449. Electronic absorption spectrum (MeCN) km”, nm (E): 285 sh (~10,100), 253 (36,600), 216 (114,600). 1H NMR (CD30N) 5 ppm: 3.46 (s, 18H, o-OCH3), 3.78 (s, 9H, p-OCH3), 6.08 (d, 4JP-H = 4.1 Hz, 6H, m-H). 31P (CD30N) 5 ppm: + 10.8 (s). 130 (CD30N) 5 ppm: 109.66 (d, 1Jp,C = 121.6 Hz, Cl), 164.40 ((1, 2Jp,C = 1.3, 02,6), 92.04 (d, 3JP-C = 6.9, 03,5), 164.00 (d, 4JP-C = 1.1, 04), 56.42 (s, o-OCH3), 55.98 (s, p-OCH3). Cyclic voltammetry (0.1 M TBABF4/CH30N, vs Ag/AgCl): Ep,c(irrev.)1 = - 0.65 V, Ep,a(chemical) = - 0.09 V, sz = -1.29 V, EMt = + 1.41 V. Mass spectrum (El), m/z: 548 (TMPP=O), 517, 367, 335, 374, 197, 181, 167, 151, 136, 121. 36 (3) Preparation of Ni(CO)3TMPP The compound Ni(CO)3TMPP was prepared in situ by a modification of the method described by Tolman for the preparation of other Ni(CO)3PR3 complexes.7 An amount of Ni(CO)4 (100 11L, 0.77 mmol) was syringed into a flask containing 15.4 mL of a 0.05 M solution of TMPP in benzene. The reaction was shaken briefly and allowed to stand for 15 min. The solution was then diluted with fresh benzene. The infrared spectrum was obtained by removing a small sample and transferring it to a 0.1 mm CaF2 solution IR cell: v(CO) (cm-1); 2048 (A1) and 1963 (E). B. X-ray Crystallography The structure of tris(2,4,6-trimethoxyphenyl)phosphine (1) and tris(2,4,6-trimethoxyphenyl)phosphine oxide (2) were determined by application of general procedures which have been fully described elsewhere.8 Geometric and intensity data were collected on a Rigaku AFCGS diffractometer with graphite-monochromated MoKa Ova = 0.71069 A) radiation and were corrected for Lorentz and polarization effects. All calculations were performed with VAX computers on a cluster network within the Department of Chemistry at Michigan State University using the Texsan software package of the Molecular Structure Corporation.9 Important crystallographic parameters for the structures of 1 and 2 are summarized in Table 2. (l) TMPP (i) Data Collection and Reduction. Large single crystals of TMPP (1) were obtained by slow evaporation of a benzene solution of TMPP in air. A suitable single crystal, with the approximate dimensions of 0.49 x 0.47 x Table 2. Summary of Crystallographic Data for TMPP (1) and TMPP=O - 37 2H20 (2) Cnmnmmd 1 L Formula P09027H33 P0120.271'137 Formula weight 532.53 548.53 Space group Pn (#7) P-1 (#2) a, A 8.092 (4) 12.768 (4) b, A 24.871 (4) 13.506 (5) c, A 13.455 (3) 9.927 (4) 01, deg 90 109.56 (3) [3, deg 96.83 (3) 109.96 (3) 7, deg 90 98.99 (3) v, A3 2689 (1) 1444 (2) Z 4 2 dcalc, g/cm3 1.315 L262 0 (Mo Koc), cm-1 1.47 1.41 Temperature, °C 23 °C -80°0 Trans. factors, max, min. 1.00, 0.95 1.00, 0.95 R8 0.040 0.045 wa 0.038 0.051 Quality-of-fit indicatorc 2.08 2.65 aR=2 I IFOI- chH/ZIFOI bRw = [2w | F0 l— I Fc I)2/2w IF, I211/2; w = 1/62( IFo I) CQuality of fit = [XW( lFo I - [Fe |)2/(Nobs-Nparameters)]l/2 38 0.31 mm3, was cleaved from a larger crystal and mounted on the tip of a glass fiber with epoxy cement. Cell constants were obtained from a least-squares refinement using 24 carefully centered reflections in the range 21 < 20 < 32°, and were found to correspond to a monoclinic cell with the following cell constants: a = 8.092 (4) A, b = 24.871 (4) A, c = 13.455 (3) A, a = y = 90°, (3 = 96.83 (3)°, V = 2689 (1) A3. Data were collected at 23i2°0, by using the (0— scan method, in the range 4 .<_ 20 S 47°. Weak reflections, those with F < 10.06(F), were rescanned at a maximum of 2 rescans and the counts were accumulated to ensure good counting statistics. A total of 4403 reflections were collected. Equivalent reflections were averaged (Rmerge = 2.6%) to yield a total of 4088 unique data and 3035 data with F02 2 36(F0)2. Periodic measurement of three representative reflections at regular intervals showed no loss of diffraction intensity had occurred during data collection. An empirical absorption correction was applied on the basis of azimuthal scans of 3 reflections with x near 90° and resulted in transmission factors ranging from 0.95 to 1.00. (ii) Structure Solution and Refinement. The positions of all non- hydrogen atoms were obtained by direct-methods10 and refined by successive full-matrix least-squares cycles. All non-hydrogen were refined with anisotropic thermal parameters. Hydrogen atoms were treated as fixed contributors at idealized positions and were not refined. Final least squares refinement of 665 parameters resulted in residuals of R = 0.040 and Rw = 0.038 and a goodness-of-fit = 2.08. Refinement of both enantiomorphs established the correct one at the 97.5% confidence level by application of the Hamilton Significance test.11 A final difference Fourier map revealed no peaks above 0.42 e'/A3. 39 (2) TMPP=O - 2 H20 (i) Data Collection and Reduction. A large crop of colorless crystals were grown in air by slow evaporation of a benzene solution of TMPP=O. A single crystal of approximate dimensions 0.36 x 0.13 x 0.10 mm3 was taken up with silicon grease on the end of a glass fiber and placed in a cold N2 stream. Least squares refinement of 21 orientation reflections in the range 20 < 20 < 38° resulted in cell constants consistent with a triclinic cell. Intensity measurements were performed at -80 i 2°C using the (1)-20 scan technique. Routine measurement of three check reflections at regular intervals throughout data collection revealed that the crystal had experienced only a slight decay (1.6%) in diffraction intensity. Nonetheless, a linear decay correction was applied to compensate for the observed loss. In addition, an empirical absorption correction was applied based on azimuthal scans of 3 reflections with Eulerian angle 1 near 90°. A total of 4476 data were collected in the range of 7.36 S 20 S 47 °. After averaging equivalent reflections (Rmerge = 2.1%), there remained 4254 unique data of which 3045 were in the category F02 2 36(F0)2. (ii) Structure Solution and Refinement. The positions of all non- hydrogen atoms were obtained by application of the direct methods programs MITHRIL and DIRDIF.12 Hydrogen atoms of the phosphine oxide were placed at calculated positions, while the positions of the hydrogen atoms of the interstitial H20 molecules were located from difference Fourier maps. All hydrogen atoms were treated as fixed contributors to the structure factor calculation and were not refined. Final least-squares refinement of 361 parameters gave residuals of R = 0.045 and RW = 0.051 and a goodness-of-fit of 2.65. The highest peak remaining in the Fourier difference map was 0.22 e'/A3. 40 3. Results and Discussion A. Synthesis and Characterization The tertiary phosphine compound tris(2,4,6-trimethoxyphenyl)- phosphine (TMPP) (1)can be prepared in good yield (60-7 0%) by lithiation of 1,3,5-trimethoxybenzene followed by coupling with triphenylphosphite at 0°C (eq 8). The phosphine was isolated as a white solid after recrystallization from hot ethanol. Alternatively, TMPP can be prepared in an unusual one- pot synthesis following a Soviet literature procedure.131 This method entails OMe OMe OMe "BuLi P(oph)3 W W P @- OMe (8) MeO OMe 0°C MeO OMe 12h OMe 3 12h refluxing 1,3,5-trimethoxybenzene with zinc chloride in neat P013 for 8 hours (eq 9). Yields are comparable to those obtained from the lithiated arene. Both procedures are currently used in our laboratories. OMe OMe O P013 06H“ Zn012°©30M 211012 MeO OMe A, 8h OMe (9) 2) H20 P Q OM 3) Na2SO4 OMe 3 Despite the known sensitivity of many phosphine lingands towards aerial oxidation, TMPP is air-stable both in solution and in the solid state. The phosphine is highly soluble in acetone and acetonitrile and moderately 41 soluble in THF, MeOH and EtOH. The molecule exhibits limited solubility in both benzene and toluene, is relatively insoluble in diethyl ether, and is completely insoluble in water and hydrocarbon solvents. Although soluble in methylene chloride, TMPP readily decomposes (t1/2 < 15 min) to form the chloromethyl phosphonium salt [(06H2(0Me)3)3P-CH201+][01'] (eq 10).2 MeO MeO + CH2C12 _ MeO P ——-> MeO P-CH2C1 Cl (10) MeO 3 MeO 3 The infrared spectrum of TMPP reveals a number of vibrations between 400 and 1600 cm'1 that render the presence of TMPP in a sample readily apparent. The 1H NMR spectrum of TMPP in MeCN exhibits three resonances: 3.45 (s,18H, o-OMe); 3.75(s, 9H, p-OMe); 6.07(d, 4Jp_H = 2.51 Hz, 6H, m-H) corresponding to the meta-ring protons, and the ortho- and para- methoxy protons, respectively. The appearance of only 3 resonances indicates that fast rotation about the P'Cipso bond is equilibrating all-exchangeable meta-ring and methoxy protons. This rotation is quite facile, in so much as the 1H NMR spectrum remains unchanged down to -100°C in d8-toluene. The 31P{1H} NMR spectrum in CD3CN exhibits a single resonance at -66.3 ppm. The extreme upfield chemical shift is attributed to an unusually large 7 effect caused by the presence of the ortho-methoxy substituents on the aryl rings of TMPP.12 The 7 effect is believed to arise from the additional shielding provided by the proximity of an ortho substituent to the phosphorus lone pair. The effect is lost upon quaternarization of the phosphine and the 31P chemical shifts of the corresponding phosphonium salts are similar to non-ortho substituted phosphonium salts.13 42 In addition to the spectroscopic properties of TMPP, we are also interested in its nucleophilicity. In order to ascertain the donating capabilities of TMPP to metal centers, we measured the A1 mode of the carbonyl stretching vibration of the nickel complex Ni(CO)3TMPP. The complex was prepared in situ by the method described by Tolman to measure the donor properties of a host of other phosphines.7 The IR spectrum of Ni(CO)3TMPP measured in benzene registers 2 00 bands, 2047.9 cm'1 and 1962.7 cm'1 which are assigned to the A1 and E stretching modes respectively. 14 A comparison of the A1 stretching mode of TMPP with those of other phosphines is shown in Figure 4.15 The stretching frequency is the lowest reported for a phosphine complex of Ni(CO)3 and is indicative of the highly donating nature of TMPP. This result agrees well with the reported pKa value for TMPP (pKa = 11.02)3 and further supports that TMPP is one of the most strongly donating tertiary phosphines. The oxide derivative of TMPP was prepared by refluxing an acetone solution of TMPP with aqueous H202 and was isolated as a white crystalline solid in high yield (87%) (eq 11). The TMPP oxide molecule is air-stable and has similar solubility properties as those of the parent phosphine.Attempts to prepare TMPP=O by reaction of TMPP and Me3NO under refluxing conditions met with failure, yielding only unreacted phosphine. Other work has shown that TMPP oxide can be prepared by catalytic oxidation of TMPP with 02 in the presence of FeCl3.16 TMPP + 30% H202(aQ) $7 TMPP=O (11) 1 2 The infrared spectrum of TMPP=O displays a number of vibrations between 400 and 1600 cm'1 that are slightly shifted from those found for the 43 .TMPP B , P 180° - ' U 3 . (Measi)3P : 0 < 160 - 2 O . Ph3P 0 140° — .MePhZP Et3P . . (PhO)3P o _ ° MeZPhP 120 MeaP . l l 01(MeO)3P 2050 2060 2070 2080 (cm") vA1(CO) for LNi(c0)3 Figure 4. Plot of the cone angle versus the v(CO)A1 stretch for various Ni(CO)3(PR3) complexes. 44 parent phosphine. The v(P-O) band is obscured by the other vibrations of TMPP in this region and is not discernible. The purity of the solid product was confirmed by 1H and 31P NMR spectroscopy. The 1H NMR spectrum of TMPP=O in 0D3CN exhibits three resonances at 5 ppm, 3.46 (s, 18H, o-Me), 3.78 (s, 9H, p-OMe) and 6.08 (d, 4JP-H = 4.1 Hz, 6H, m-H). As was observed for TMPP, the appearance of only three resonances suggests that the aryl groups are free to rotate about the P'Cipso bond. In spite of the change in phosphorus oxidation state, these resonances are only slightly shifted from the free phosphine. In sharp contrast, the 31P chemical shift difference between TMPP and TMPP=O is over 75 ppm, as the 31P displays a single resonance at +10.8 ppm. This dramatic chemical shift difference is a result of deshielding of the phosphorus nucleus upon oxidation from +3 to +5 and illustrates the sensitivity of 31P NMR spectroscopy to changes in phosphorus oxidation state. B. Molecular Structures (1) TMPP Single crystals of TMPP were grown by slow evaporation from a benzene solution. The compound was found to crystallize in the acentric space group P2n (#7), resulting in the presence of two independent TMPP molecules per asymmetric unit. Refinement of the structure in the centric space group P21/n (#13) proved unsatisfactory as evidenced by abnormally high residuals. Therefore, the final refinement was performed in the acentric group. Further evidence for the acentric space group was provided by the computer program MISSYM,17 which showed that the two independent molecules are not symmetry related. ORTEP drawings for both independent molecules are depicted in Figure 5. Pertinent bond distances and angles are Summarized in Table 3. 45 As expected, the aryl rings of the phosphine adopt a pseudo-propeller arrangement about the phosphorus atom. However, the angles that the aryl rings adopt with respect to the plane described by the ipso-carbon atoms vary considerably from ideality, ranging in value from 36.61° to 74.12° (Table 4). The average P'Cipso bond length, 1.842 A, is slightly longer than that reported for triphenylphosphine (P-Cave = 1.828 A),13 but comparable to that reported for the structure of tris(2,6-dimethoxyphenyl)phosphine (P-Cave = 1.844 A)19 and trimesitylphosphine (P'Cave = 1.837 A).20 As expected, the average C-P-C bond angle of TMPP (C-P-Cawe = 105.2°) is larger than that of triphenylphosphine (C-P-Cave = 103.0°), yet smaller than trimesitylphosphine (C-P-Cawe = 109.7°). Although the presence of the ortho-methoxy groups contributes to the deformation of the C-P-C bond angle, the steric interactions between these groups are not nearly as great as those observed for the analogous methyl substituted phosphines tris(2,6-dimethylphenyl)phosphine and trimesitylphosphine. As a result, the cone angle of TMPP is not as large as those observed for the methyl substituted analogs. Just as was observed in the structure of the 2,6-dimethoxy substituted phosphine, the ortho-methoxy groups of TMPP are bent slightly towards the phosphorus lone pair. For example, the O-C-C angle, O(3)—C(6)-C(5) is 122.3(8) while the O-C-C angle O(3)-C(6)-C(1) is 116.0(8), indicating a bending towards the P-C bond. This effect is opposite to that observed for trimesitylphosphine, in which the methyl groups bend away from the P-C bond. The methoxy groups of phosphine are approximately in the plane of the phenyl ring with the methoxy group torsion angles relative to the phenyl rings varying from 032° to 21.42°. NEE mo 318538 use :25 bfioEnmuMozaamio 23 on... no macs—Sawmemouvmmemam .m 959.,— 46 47 Table 3. Selected bond distances (A) and angles (deg) for TMPP (1). Atom 17 Atom 2 Distance Atom 1 Atom 2 Distance P(1) C(l) 1.823(8) P(2) C(28) 1.835(7) P(1) C(10) 1.847(8) P(2) C(37) 1.850(7) P(1) C(19) 1.842(7) P(2) C(46) 1.853(7) 0(1) C(2) 1.377(9) 0(10) C(29) 1.342(8) 0(1) C(7) 143(1) 0( 10) C(34) 1.38(1) C(2) C(4) 140(1) 0(11) C(31) 1.368(8) 0(2) C(8) 135(1) 0(11) C(35) 142(1) 0(3) C(6) 135(1) 0(12) C(33) 1.372(9) 0(3) C(9) 134(1) 0(12) C(36) 1.442(9) Atom 1 Atom 2 Atom 3 Bond angle C(1) P(1) C(10) 106.4(3) 0(1) P(1) C(19) 101.6(3) C10) P(1) C(19) 108.1(3) C(28) P(2) C(37) 102.5(3) C(28) P(2) C(46) 106.5(3) C(37) P(2) C(46) 106.2(3) 48 Table 4. Dihedral angles (deg) between the planes of the aryl rings and the plane described by Cipso for TMPP (1) and TMPP=O - 2H20, (2). TMPP (1) TMPP=O - 2 H20 (2) 0(1) - 0(6) 74.12 89.66 C(10) - C(15) 56.41 28.81 C(19) - C(24) 40.05 41.71 C(28) - C(33) 72.09 ------- C(37) - C(42) 53.55 ....... C(46) - C(51) 36.61 ....... 49 (2) TMPP=O - 2 H20 An ORTEP drawing of TMPP oxide is presented in Figure 6. A listing of selected bond distances and angles is given in Tables 5. In contrast to the propeller-like geometry of phenyl rings of TMPP, only two of the phenyl rings of TMPP=O are tilted at 28.81° and 41.71° with respect to the plane described by the ipso-carbon atoms. The third phenyl group is perpendicular to this plane and hence parallel to the P=O bond. (Table 4). A similar arrangement of the phenyl rings was observed in the structure of tris(2,6 dimethoxyphenyl) phosphine. The average P'Cipso distance (1.817A) is slightly contracted from that of the parent phosphine and the average C-P-C angle has opened up from 105.20 in TMPP to 107.8°. However, the flattening of the C-P-C angles is not as extensive as that observed for the metal bound phosphine oxide in the structure of FeCl3(O=TMPP) (C-P-Cavez 111.2[5]°).6 Not surprisingly the P-O bond distance is much shorter in the free phosphine oxide than in the coordinated ligand (Rom: 1.497(2) A vs P-oc,,,d= 1.550(7) A), but nevertheless, it is slightly longer than that observed for other documented structures of phosphine oxide ligands. One explanation for this is the presence of two hydrogen bonded water molecules in the structure of TMPP=O (vide infra). Finally, the observed torsion angles involving the methoxy groups and the phenyl rings vary from 286° to 9.66°, placing the methoxy substituents in an essentially co-planar arrangement with the phenyl rings. Interestingly, although the TMPP=O compound was crystallized from dry benzene, the structure shows that there are two H20 molecules per molecule of TMPP=O. Obviously, these were carried along in the crude product prepared in 30% H202 solution. The location of the H20 molecules 50 .ONEN . Dung mo Ear—MSU mmfimo .@ 953% km: NS Figure 7. Packing diagram of TMPP=O - 2H20. 52 Table 5. Selected bond distances (A) and angles (deg) for TMPP=O . 2H20 (2). Atom 1 Atom 2 Distance Atom 1 Atom 2 Distance P(1) O(10) 1.497(2) 0(1) 0(2) 1.367(4) P(1) 0(1) 1.813(4) 0(1) 0(7) 1.421(4) P(1) C(10) 1.810(3) 0(2) 0(4) 1.375(4) P(1) C(19) 1.828(4) 0(2) 0(8) 1.433(5) 0(1) 0(2) 1.401(5) 0(3) 0(6) 1.369(4) 0(1) 0(6) 1.397(5) 0(3) 0(9) 1.398(5) C(10) C(11) 1.394(5) 0(4) C(11) 1.365(4) C(10) C(15) 1.401(5) 0(4) C(16) 1.421(4) C(19) C(20) 1.403(5) 0(5) C(13) 1.373(4) C(19) C(24) 1.410(4) 0(5) C(17) 1.432(5) Atom 1 Atom 2 Atom 3 Angle O(10) P(1) C(1) 114.2(2) O(10) P(1) C(10) 107.9(1) O(10) P(1) C(19) 111.5(2) 0(1) P(1) C(10) 110.0(2) 0(1) P(1) C(19) 101.5(2) C(10) P(1) C(19) 11.8(2) 53 suggests that they are both hydrogen bonded to the terminal oxygen atom of the phosphine oxide. Their positions relative to the phosphine is clearly seen in the packing diagram (Figure 7). The hydrogen atoms of the water molecules were located in the Fourier difference map but were treated only as fixed contributors to the structure factor calculation and their positions were not refined. Hydrogen atoms, H(34) and H(36), lie at a distance of 1.82 A and 1.86 A from O(10) and are clearly interacting with the oxygen lone pairs. The resulting P-O---H bond angles are approximately 124.5° and 141.7°. Apparently, the occurrence of a phosphine oxide involved in two hydrogen bonds is rare, as only three other cases have been reported.21 In each of the reported structures, the P-O bond is slightly longer than that of the non hydrogen bonded phosphine oxide. 10. 11. 12. 13. 54 List of References (a) Protopopov, I. S.; Kraft, M. Y. Zhurnal Obshchei Khimii 1963, 33, 3050. (b) Protopopov, I. S.; Kraft, M. Y. Med. Prom. SSSR, 1959, 13, 5; Chem. Abstr. 1960, 54, 10914c. (a) Wada, M.; Higashizaki, S.; J. Chem. Soc., Chem. Commun. 1984, 482. (b) Wada, M.; Higashizaki, S.; Tsuboi, A. J. Chem. Res. 1985, (S), 38; (M), 0467. Yamashoji, Y.; Matsushita, T.; Wada, M.; Shono, T. Chem. Lett. 1988, 43. A preliminary structural report of TMPP has appeared: Wada, M. J. Syn Org. Chem. Jpn. 1986, 957. For examples of phosphine oxidation promoted by platinum group metals see: (a) Sen, A.; Halpern, J. J. Am. Chem. Soc. 1977, 99, 8337. (b) Van Vungt, B.; Koole, N.; Drenth, W.; Kuijpers, F. P. Rec. Trav. Chim., Pays-Bus 1973, 92, 1321. (c) Takao, K.; Fujiwara, Y.; Imanaka, T.; Teranishi, S. Bull. Chem. Soc. Jpn. 1970,43, 1153. (d) Wilke, G.; Schott, H.; Heinbach, P. Angew. Chem. Int. Ed. 1967, 6, 92. (a) Dunbar, K. R.; Haefner, S. 0.;1990, 9, 1695. (b) Dunbar, K. R.; Quillevéré, A. Polyhedron in press. Tolman, C. A. J. Am. Chem. Soc. 1970, 92, 2953. (a) Bino, A.; Cotton, F. A.; Fanwick, P. E. Inorg. Chem. 1979, 18, 3558. (b) Cotton, F. A.; Frenz, B. A.; Deganello, G.; Shaver, A. J. Organomet. Chem. 1973,50, 227. TEXSAN - TEXRAY Structure Analysis Package, Molecular Structure Corporation, 1985. (a) MITHRIL: Integrated Direct Methods Computer Program, Gilmore, C. J. J. Appl. Cryst. 1984, 17, 42-46. (h) DIRDIF: Direct Methods for Difference Structure, An Automatic Procedure for Phase Extension; Refinement of Difference Structure Factors. Beurskens, R. T. Technical Report 1984. Hamilton, W. C. Acta Cryst. 1965, 18, 502. (a) Grim, S. O.; Yankowsky, A. W. Phosphorus and Sulfur 1977,3, 191. (b) Quin, L. D.; Breen, J. J. Org. Mag. Resonance 1973,5, 17 . Grim, S. O.; Yankowsky, A. W. J. Org. Chem. 1977,42, 1236. 14. 15. 16. 17. 18. 19. 20. 21. 55 Band assignment is based on the work of (a) Stromeier, W.; Muller, F. J. Chem. Ber. 1967, 100, 2812. (b) Bigorgne, M. Bull. Soc. Chim. Fr. 1960, 1986. Tolman, C. A. Chem. Rev. 1977, 77, 313. Dunbar, K R.; Quillevéré, A. submitted for publication in Polyhedron. Le Page, Y. J. Appl. Cryst. 1988, 21 , 983. Daly, J. J. J. Chem. Soc. 1964, 3799. Livant, P.; Sun, Y. J .; Webb, T. R. Acta Cryst. 1991, C47, 1003. Blount, J. F.; Maryanoff, C. A.; Mislow, K. Tetrahed. Lett . 1975, 913. (a) Llamas-Saiz, A. L.; Foces-Foces, C.; Elguero, J.; Molina, P.; Alajarin, M.; Vidal, A. J. Chem. Soc., Chem. Commun. 199], 1694. (b) Jones, R.; Warrens, C. P.; Williams, D. J.; Woollins, J. D. J. Chem. Soc., Dalton Trans. 1987, 907. (c) Antipin, M. Yu.; Akhmedov, A. I.; Struchkov, Yu. T.; Matrosov, E. I.; Kabachnik, M. I. Zh. Strukt. Khim. 1983,24, 888. CHAPTER III ISOLATION, CHARACTERIZATION AND REDOX CHEMISTRY OF THE MONONUCLEAR Rh(II) COMPLEX [Rh(n3-TMPP)2][BF4]2 56 57 1. Introduction This chapter addresses the synthesis, solution properties, redox chemistry, and solid-state structure of a remarkable rhodium(II) phosphine complex. These findings are of considerable interest in the general context of small molecule binding to paramagnetic metal centers, which has been well investigated in the case of 00(II), but for which parallel studies are essentially non-existent for Rh(II).1 The reason for this is undoubtedly the paucity of stable mononuclear complexes Rh(II) that have been established as authentic metal-based radicals.2 The uncertainty as to the location of the unpaired electron is especially high in the reported Rh(II) compounds with non-innocent sulfur- and nitrogen-based ligands.3 Only recently, through the work of Wayland and co-workers,4 has the tremendous potential of Rh(II) metallo-radical systems become apparent. Our entry into mononuclear Rh(II) chemistry came about by a less traditional method than previous approaches, but it is one that holds great promise for preparing many previously unknown rhodium compounds. Typically mononuclear Rh(II) species are prepared by either reduction of hydrated Rh(III) halides in alcohol5 or by oxidation of Rh(I) complexes“,7 Often such routes yield impure products that are inevitably contaminated with diamagnetic species. In contrast, there have been few reports of paramagnetic Rh(II) species being prepared from dinuclear Rh(II) complexes.8 This is, in part, due to the lack of suitable dinuclear starting materials. Currently, a major focus of our research group is the development of solvated metal systems, particularly dinuclear species, as synthetic precursors for radical mononuclear complexes.9 To this end we are exploring the chemistry of the solvated Rh211,11 complex [Rh2(MeCN)10][BF4]4.10 The use of solvated metal systems in conjunction with TMPP have produced a 58 number of homoleptic phosphine complexes. For example, acetonitrile complexes of 002+ and Ni2+ react with TMPP to give neutral bis-phosphino- phenoxide compounds.11 More significantly, the reaction of [Rh2(MeCN)10][BF4]4 with TMPP has resulted in the first structurally characterized six-coordinate Rh(II) complex without a metal-metal bond. Herein, the synthesis, spectroscopy, structural characterization, and redox chemistry of the remarkably stable Rh(II) complex [Rh(n3-TMPP)2][BF4]2 (3) is presented. 2. Experimental A. Synthesis All reactions were carried out under an argon atmosphere by the use of standard Schlenk-line techniques unless otherwise stated. The solvated dirhodium complex, [Rh2(Me0N)1o][BF4]4, was prepared as described in the literature.12 Tris(2,4,6-trimethoxyphenyl)phosphine (TMPP) (1) was prepared by the reaction of triphenylphosphite with thevlithium salt of trimethoxybenzene as described in chapter II. [szFe][BF4] was prepared by oxidation of szFe with hydrofluoroboric acid in the presence of p- benzoquinone.13 The reagents NOBF4, NOPFe, and szCo were purchased from Strem Chemicals and used without further purification. (1) Preparation of [Rh(n3-TMPP)2][BF4]2 (3) An amount of [Rh2(MeCN)10][BF4]4 (0.650 g, 0.675 mmol) was dissolved in 10 mL of MeCN and cooled to -15°C with a dry ice/ethylene glycol bath. The reaction was covered with aluminum foil to avoid photochemical generation of Rh(I) and Rh(III) species.14 A solution of TMPP (1.437 g, 2.698 mmol) in 20 mL of MeCN was added dropwise over a 15 minute period. The resulting purple solution was stirred for 1 h at -15°C and then evaporated to 59 a residue. The crude purple solid was washed several times with copious amounts of THF. It was then redissolved in 50 mL of CH2012, filtered, and treated with THF (approximately 10-15 mL). The volume of the solution was reduced with the use of a rotary evaporator to approximately 10 mL to yield a purple crystalline solid. The solid was filtered, washed with 4 x 10 mL of THF, followed by 4 x 10 mL of diethyl ether, and dried in vacuo.; yield 1.647 g (91%). Anal. Calcd for C55H66F8P2018B2Rh: C, 48.34; H, 4.96. Found: C, 48.56; H, 4.86. IR (Nujol, CsI) cm'l: 1599 vs, 1585 vs, 1412 s, 1380 m, 1230 vs, 1207 vs, 1190 m, 1161 vs, 1135 s, 1110 s, 1049 vs/br, 950 m, 920 w, 908 m, 820 m, 785 w, 720 w, 700 w, 680 w, 673 w, 637 w, 520 w, 490 m, 450 w, 435 w, 415 w. Electronic absorption spectrum (0H2012) )tmax, nm (e): 537 (2050), 420 sh, 329 (13,400), 298 (17,500), 233 (62,400). 1H NMR (CDzClz) 5 ppm: Broad. 31P NMR (CD2012) 5 ppm: not observed. (2) Preparation of [Rh(n3-TMPP)2][BF4]3 (4) (i) Oxidation of [Rh(n3-TMPP)2][BF4]2 with NOBF4. A quantity of [Rh(n3-TMPP)2][BF4]2 (0.200 g, 0.14 mmol) and NOBF4( 0.016 g, 0.14 mmol) was dissolved in 5 mL of MeCN which resulted in the immediate formation of a deep red solution. The reaction was stirred at -40 °C for 45 min with periodic pumping to remove the evolved N0 gas, after which time 20 mL of diethyl ether was added to precipitate the product. The red solid was collected by filtration under argon, washed with 3 x 5 mL of diethyl ether and dried in vacuo; yield, 0.085 g (80%). Due to thermal instability, solid and solution forms of 4 must be stored anaerobically at -20°C. Anal. Calcd for C54H66F12P2018B3Rh: C, 45.40; H, 4.66. Found: C, 44.39; H, 5.14. Electronic absorption spectrum (CH2012) Amax, nm (6): 363 (22,000), 260 sh, 245 (65,000). 1H NMR (CD2012) 5 ppm: -OCH3, 2.93 (s, 6H), 3.57 (s, 6H, ), 3.59 (s, 6H), 3.64 (s, 6H), 3.90 (s, 6H), 3.92 (s, 6H), 3.97 (s, 6H), 4.19 (s, 6H), 60 4.69 (s, 6H); m-H, 5.73 (dd, 2H), 6.05 (dd, 2H), 6.18 (dd, 2H), 6.32 (dd, 2H), 6.50 (dd, 2H), 6.68 (dd, 2H). 31P NMR (CD2012) 5 ppm: 37.4 (d, lJRh_P= 107 Hz). (ii) Oxidation of [Rh(n3-TMPP)2][BF4]2 With [Cp2Fe][BF4]. A solution of [Rh(n3-TMPP)2][BF4]2 (0.100 g, 0.07 mmol) and [Cp2Fe][BF4] (0.014 g, 0.07 mmol) in 5 mL of 0H2012 was stirred at 40°C for 30 min. The red solution was treated with 20 mL of diethyl ether to precipitate a solid which was collected by filtration, washed with 2 x 10 mL of diethyl ether to remove unreacted ferrocene and finally dried in vacuo; yield, 0.091 g (85%). (iii) Oxidation of [Rh(n3-TMPP)2] [BF4J2 with HBF4 . Et20. An amount of [Rh(n3—TMPP)2][BF4]2 (3) (0.015 g, 0.011 mmol) was dissolved in ~ 0.7 mL of CD3CN. The purple solution was then transferred to a NMR tube and ~ 0.1 mL of HBF4 o EtZO. Within 12 h, the solution colored had changed from purple to red. A 31P NMR spectrum of the solution at this time revealed the presence of [RhHI(n3-TMPP)2][BF4]3 (4). (3) Demethylation of [Rh(n3-TMPP)2][BF4]3: Formation of [Rh(n3- TNIPP)(C6H2(0Me)20P{06H2(0Me)3}2)] [BF412 (5) (i) Solution A quantity of [Rh(n3-TMPP)2][BF413 (4) (0.150 g, 0.11 mmol) was dissolved in 10 mL of acetone to give a red solution which was stirred overnight at r.t. The pale orange solution was filtered through Celite and treated with 20 mL of diethyl ether to produce an orange solid which was collected by suction filtration, washed with 2 x 5 mL of diethyl ether, and dried under a reduced pressure; yield, 0.085 g (61%). Anal. Calcd for 053H63F8P201882Rh: C, 47.98; H, 4.79. Found: C, 47.28; H, 5.21. Electronic absorption spectrum (CH2012) Amax, nm (e): 329 (25,000) 260 sh, 242 sh. 1H NMR (CD2012) 5 ppm: -OCH3, 2.94 (s, 3H), 3.06 (s, 3H), 3.32 (s, 3H), 3.41 (s, 3H), 3.46 (s, 3H), 3.52 (s, 3H), 3.53 (s, 3H), 3.55 (s, 3H), 3.68 (s, 3H), 3.84 (s, 61 3H), 3.85 (s, 6H), 3.90 (s, 3H), 3.92 (s, 3H), 4.16 (s, 3H), 4.33 (s, 3H), 4.51 (s, 3H); m-H, 5.60 (mult, 3H), 5.84 (dd, 1H), 5.89 (dd, 1H), 5.98 (dd, 1H), 6.11 (dd, 1H), 6.19 (dd, 1H), 6.23 (dd, 1H), 6.36 (dd, 1H), 6.53 (dd, 1H), 6.96 (dd, 1H). 31P NMR (CD2012) 5 ppm: + 31.5 (dd, 1JRh_p= 140 Hz, 2JPA_PB= 13.7 Hz), + 37.9 (dd, 1JRh_P= 139 Hz, 2JPA_PB= 13.7 Hz). (ii) Solid State. An amount of [Rh(n3-TMPP)2][BF4]3 (4) (0.015 g, 0.010 mmol) was placed in a Schlenk tube and heated to 50°C for 36 hrs in an oil bath, during which time the color of the solid changed from deep red to pale orange. After cooling to room temperature, the 1H NMR spectrum of the solid was recorded and compared to an authentic sample of 5 prepared as in (1). Conversion of 4 to 5 was quantitative based on this result. (iii) Reaction of [Rh(n3-TMPP)2][BF4]3 (4) with TMPP. A quantity of [Rh(n3-TMPP)2][BF4]3 (4) (0.015 g, 0.010 mmol) and TMPP (0.006 g, 0.011 mmol) was dissolved in ~ 0.7 mL of CD3CN. The solution color immediately went from red to orange. The solution was pippetted into a NMR tube and the conversion of 4 to 5 was confirmed by 1H NMR spectroscopy. (iv) Reaction of [Rh(n3-TMPP)2][BF4]3 (4) with [(Bun)4N] [I]. A mixture of [Rh(n3-TMPP)2][BF4]3 (4) (0.015 g, 0.010 mmol) and tetra-butyl ammonium iodide (0.004 g, 0.011 mmol) were dissolved in ~ 0.07 mL of d6- acetone. The solution color changed from red to orange within 5 min. The solution was transferred to a NMR tube in order to monitor the reaction progress by 1H and 31P NMR spectroscopy. Conversion of 4 to cis-[RhHI(n3- TMPP)(n3-TMPP-O)][BF4]2 (5) was found to be quantitative by 31P NMR spectroscopy. The presence of 0H3I (5 = 2.16 ppm), presumably formed as a by-product, was detected by 1H NMR. 62 (4) Chemical Reduction of [Rh(n3-TNIPP)2][BF4]2 In a typical reaction, a mixture of [Rh(n3-TMPP)2][BF4]2 (0.100 g, 0.07 5 mmol) and szCo (0.014 g, 0.075 mmol) was dissolved in acetonitrile (5 mL). The resulting purple solution was stirred for 5 h at r.t. During this period the solution color transformed from purple to red-brown. The solvent was removed under vacuum and the resulting residue was taken up in THF (5-10 mL) leaving behind undissolved [Cp2Co][BF4]. The red-brown solution was filtered through Celite and subsequently pumped to dryness. Finally, ' the brown residue was washed with diethyl ether (2 x 10 mL) and dried under reduced pressure. 1H NMR (d6-acetone) 5 ppm: -OCH3: 3.46 (s, 3H), 3.47 (s, 6H), 3.52 (s, 18H), 3.67 (s, 3H), 3.80 (s, 12H), 3.86 (s, 6H), 3.90 (s, 3H); m-H: 4.83 (mult, 1H), 5.18 (mult, 1H), 5.87 (dd, 1H), 6.08 (t, 4Jp_H = 4JHa-Hb = 2.1 Hz, 1H), 6.17 (d, 4JP_H= 2.7 Hz, 4H), 6.26 (d, 4Jp_H= 3.6 Hz, 2H), 6.40 (d, 4Jp_H= 3.0 Hz, 2H). 31P NMR (CD30N) 5 ppm: -8.9 (dd, 1JRh_p= 131.8 Hz, 2JpA_PB= 392.4 Hz), -134 (dd, 1JRh_p= 120.9 Hz, 2JPA_pB= 392.4 Hz). Samples were contaminated with a small amount of TMPP-CH3+, 1H NMR (d6-acetone) 5 ppm: 2.47 (d, 2Jp,H= 15 Hz, 3H, P-CH3), (3.66 (s, 18H, o-OME), 3.84 (s, 9H, p-OMe), 6.34 (d, 4JP-H= 4.8, 6H, m-H); and [Cp200]+: 1H NMR (d6-acetone) 5 ppm: 5.92 (s, 10H, Cp-H). Similar results were also obtained when acetone was used as the reaction solvent. (5) Preparation of [Rh(TMPP)2][PF6]2 In an Erlenmeyer flask, a quantity of [Rh(TMPP)2][BF4]2 (0.100 g, 0.975 mmol) and NaPF6 (0.125 g, 0.750 mmol) was dissolved in 5 mL of acetone. The solution was stirred for 15 min, filtered and an additional amount of NaPF6 was added (0.125 g, 0.750 mmol). Again, the solution was stirred for 15 minutes and filtered. The solvent was evaporated and the purple solid was redissolved in 20 mL of CH2012. After filtration of the 63 solution to remove the undissolved sodium salts, the solution was treated with 10 mL of THF and concentrated on a rotary evaporator to approximately 5 mL. The resulting purple crystalline solid was collected by suction filtration, washed with cold THF (2 x 3 mL), diethyl ether (2 x 5 mL) and finally dried under reduced pressure; yield, 0.085 g (78%). B. X-ray Crystallography The structures of the complexes [Rh(n3-TMPP)2][BF4]2 (3) and [Rh(n3- TMPP)2][BF4]3 (4) were determined by application of general procedures which have been fully described elsewhere.15 Geometric and intensity data were collected on a Nicolet P3/F diffractometer with graphite monochromated MoKa (15 = 0.71073 A) radiation and were corrected for Lorentz and polarization effects. All calculations were performed on a VAXSTATION 2000 computer. Data reduction and the initial refinement were performed using the programs from the Enraf-Nonius Structure Determination Package (SDP).16 The modeling of the [BF4]' ion and final refinement for 4 were carried out with the use of the Texsan crystallographic software package of Molecular Structure Corporation.17 (1) [Rh(n3-TMPP)2][BF4]2 (3). (1) Data Collection and Reduction. Single crystals of 3 were grown by a careful layering of toluene on a solution of the compound in 0H2012. A red- purple parallelepiped with approximate dimensions 0.67 x 0.35 x 0.22 mm3 was mounted on the tip of a glass fiber with epoxy cement. Geometric and intensity data were collected at 22 i 2°C. Indexing and refinement of 16 reflections in the range 4 S 20 S 12° selected from a rotational photograph gave unit cell parameters for an orthorhombic crystal system. The cell was 64 Table 6. Summary of crystallographic data for [Rh11(n3-TMPP)2][BF4]2 (3) and [Rhmm3-TMPP)2][BF4][PF6]2 (4a). 3 4a Formula RhP2F8018054BzH66 RhP4F16018Cs4BiH66 Formula weight 1341.5 1544.68 Space group Pbcn Pcca a, A 15.938(5) 21.205(7) b, A 17.916(7) 11.694(6) c, A 21.015(8) 2910(2) 0, deg 90 90 B, deg 90 90 7, deg 90 90 V, A3 6001(6) 7216(6) Z 4 4 dcalc, g/cm3 1.427 1.422 M, cm'1 4.20 4.17 Temperature, °C 22 i 2 -90 i 3 Trans. factors, max, min. 1.00, 0.87 1.00, 0.38 Ra 0.073 0.118 wa 0.089 0.139 fluality-of-fitc 2.89 4.57 aR=2 l IFOI- chH/ZIFOI bRw = [2w( | F0 I— In, l)2/Zw F0 |2]1/2; w = 1/0'2( lFo I) cQuality-of-fit = [EW( ”:0 I ' [ Fc I)2/(Nobs‘NparameteerU2 65 further refined by a least squares fit of 20 reflections in the range of 13 S 20 S 25°. The Laue class was determined to be mm by axial photography. A total of 3952 unique data were collected in the range 4 S 29 S 45° by using the 0-20 scan technique. Three standard reflections, measured at regular intervals every 97 reflections, decayed by 2%; a decay correction was applied to the data by using the program CHORT in SDP. After data reduction, a total of 2462 reflections remained with F02 > 30(F02). (ii) Structure Solution and Refinement. The position of the Rh atom was located by the direct methods program in SHELXS-86.18 The remaining non- hydrogen atoms were located through successive cycles of least-squares refinements and difference Fourier maps. After isotropic convergence had been achieved, an empirical absorption correction based upon the program DIFABS was applied to the data.19 In the end, refinement of 369 parameters gave residuals of R = 0.080 and Rw = 0.098 and a quality-of-fit index of 2.89. The largest shift/esd was 0.58 and the highest peak in the final difference Fourier map was 1.91 e-/A3. (2) [Rh(n 3-TMPP )2] [PF612IBF4] (1) Data Collection and Reduction. Dark red crystals of [Rh(n3- TMPP)2][PF6]2[BF4] were prepared by oxidation of [Rh(n3-TMPP)2][BF4]2 (3) with N OPF6 and were grown from a mixture of CHgCN/toluene/Etgo at -5°C in the presence of excess PF6: A crystal with the approximate dimensions 0.52 x 0.39 x 0.31mm3 was selected, taken up on the tip of a glass fiber with viscous oil and immediately placed in a cold stream of N2 at -90°C. Indexing and refinement of 13 reflections selected from a rotational photograph gave cell parameters consistent with an orthorhombic space group. After further refinement of 20 reflections in the range 13 < 20 < 25°, the Laue symmetry was determined to be m by axial photography. Data were collected using 66 the co-scan technique. Three check reflections were monitored at regular intervals throughout data collection and showed an average decay of 6%. The data were corrected for the observed decay by application of the program CHORT in SDP. After averaging equivalent reflections, 4722 unique data remained of which 2390 were observed with F02 2 30(F0)2. (ii) Structure Solution and Refinement. The position of the Rh atom was located by the direct methods program in SHELXS-86. The remaining non-hydrogen atoms were found after a series of alternating least-squares refinements and difference Fourier maps. The positional and thermal parameters of the hydrogen atoms were calculated and treated as fixed contributors to the structure factor calculation. An absorption correction was applied using the program DIFABS,19 which resulted in relative minimum and maximum transmission factors of 0.38 and 1.00. All non-hydrogen atoms of the [Rh(TMPP)2]3+ cation were refined anisotropically with the exception of C(10), C(13), C(20), C(21), C(24), and C(25) which were refined isotropically. Thermal parameters for the [PF6]' and [EMT anions were also refined isotropically. Treatment of the [BF4]' ion is described in the following section. The final refinement converged with residuals R and Rw of 0.118 and 0.139, respectively, and a quality-of-fit of 4.60. After convergence, the largest shift/esd was 0.58 and the highest peak in the difference Fourier map was 2.23 e-/A3 which is associated with the boron atom of the [BF4]- counterion. The first counterion located in the difference map occupied a general position in the cell and was determined to be a [PF6]' ion based on the initial height of the peak and the presence of six other peaks in the difference map distributed in an octahedral arrangement about the central atom. The central atom of the second counterion was found to reside on a crystallographic two-fold rotation axis and was accompanied by two other 67 atoms situated with a bond angle that approximated that of a tetrahedron. Refinement of the group as a [BF4]' ion led to a satisfactory thermal parameter for boron, but unfortunately, the fluorine atoms could not be successfully refined. Therefore the entire [BF4]' anion was treated as a rigid group at a population of 0.5 using the rigid group parameters located in the Texsan software package. The group was fixed on the two-fold axis and allowed to rotate relative to this axis. This resulted in the formation of two rigid tetrahedral [BF4]' units, at 0.5 occupancy each, that shared a central boron atom. Isotropic refinement of the [BF4]' moiety led to the formation of a pseudo-cage of eight fluorine atoms at a population of 0.5 each about the boron atom. Other attempts to refine this group as part of a PF6' ion failed, leading to an unsatisfactorily high isotropic thermal parameter (> 35) for the central atom. 3. Results The highly unusual mononuclear rhodium(II) complex, [Rh(n3- TMPP)2][BF4]2 (3) was isolated in high yield from the reaction of TMPP and [Rh2(MeCN)1o][BF4]4, and although the complex possesses extraordinary air and thermal stability for a radical species, it nonetheless exhibits a rich and varied chemistry. Chemical or electrochemical oxidation of 3 produces the d6 Rh(III) complex [Rh(n 3-TMPP)2][BF4]3 (4) with a ligand arrangement identical to that found in the parent complex. In striking contrast to the remarkable stability of the divalent complex [Rh(n3-TMPP)2][BF4]2 (3), the Rh(III) species 4 is air-sensitive and thermally unstable, eventually forming the demethylated complex Rh(III) complex, [Rh111(n3-TMPP)(n3-TMPP- 0)][BF4]2 (5), where TMPP-O = [(006H2(OMe)2P{C6H2(OMe)3)2]1'. 68 The structures of compounds 3 and 4 have been determined by single crystal X-ray diffraction methods. Crystal parameters and basic information pertaining to data collection and structure refinement are summarized in Table 6. ORTEP representations of [Rh(n3-TMPP)2][BF4]2 (3) and [Rh(n3- TMPP)2][PF6]2[BF4] are depicted in Figures 9 and 13. Selected bond distances and angles for each structure are listed in Tables 8 and 9. Full tables of positional and anisotropic thermal parameters for compounds 3 and 4 are located in the appendices. A. Preparation and Spectroscopic Properties of [Rh(n3-TMPP)2] [BF4l2 (3) Slow addition of TMPP to [Rh2(Me0N)1o][BF4]4, both dissolved in CH3CN, produces a dark red-purple solution of 3, with the reaction being complete within one hour (eq 12). The reaction is performed in the dark in order to avoid formation of Rh(I) and Rh(III) species that result from room light photolysis of [Rh2(MeCN)1o]4+ in acetonitrile.20 Free TMPP reacts with [Rh(n3-TMPP)2][BF4]2 to dealkylate one of the coordinated methoxy M N [Rh2(MeCN)10][BF4]4 + 4TMPP —%p 2 [RhII(n3-TMPP)2][BF4]2 (12) 1 3 groups. This results in the formation a new Rh(II) complex that is ligated by one neutral TMPP ligand and a phenoxy-phosphine derivative of TMPP. Therefore, in order to maximize yields of 3, the presence of excess phosphine must be avoided during the reaction. As a result, dilute solutions of TMPP are added dropwise to highly concentrated of [Rh2(MeCN)10][BF4]4 to maintain an excess of [Rh2(MeCN)10][BF4]4. Moreover, the reaction is kept at 0° C to further retard nucleophilic attack by free phosphine. Unlike 69 [RhH(n3-TMPP)2][BF4]2 (3), the demethylated Rh(II) complex is air-sensitive and readily decomposes to an intractable mixture of diamagnetic species. Details of the deliberate synthesis of this species by reaction of 3 with various nucleophiles is presented in chapter VII. Initially, [Rh(n3-TMPP)2][BF4]2 was synthesized by addition of a methanolic solution of TMPP to a suspension of [Rh2(MeCN)10][BF4]4 in MeOH. This method was attractive because the relatively low solubility of 3 in MeOH leads to the precipitation of the product. However, the yields were ' typically lower than those found for reactions carried out in MeCN; this is presumably due to the low solubility of the Rh24+ salt in MeOH which serves to keep the phosphine ligand in excess and increases the likelihood of the demethylation side-reaction. Furthermore, samples of 3 obtained from methanolic solutions were invariably contaminated by a gray solid with unusual properties. When redissolved in acetonitrile, the solid produces orange solutions that analyzed as [Rh2(MeCN)10]4+ by 1H NMR. This gray solid is believed to be a mixed-valence linear tetramer formed by one electron reduction of the dirhodium species.21 Evidence for the formation of this species is found in the photochemistry of [Rh2(MeCN)10][BF4]4 and a similar species has been prepared directly by electrochemical reduction of [Rh2(MeCN)10][BF4]4 in high ionic strength media by another member of this research group.20’21 As a final synthetic approach to preparing Rh(II) complexes of TMPP, we reacted TMPP with RhCl3'xH20 in refluxing ethanol, a method that has been widely used to prepare many of the reported Rh(II) species with bulky phosphines.5 In the present case, however, an intractable mixture of diamagnetic products was detected by 1H NMR spectroscopy. We rationalize 70 that this is largely due to the demonstrated instability of [Rh(n 3- TMPP)2][BF4]2 in the presence of free halides. The compound [Rh(n3-TMPP)2][BF4]2 is soluble in methylene chloride and acetonitrile, partially soluble in methanol, chloroform and acetone, and completely insoluble in water, tetrahydrofuran, diethyl ether and hydrocarbon solvents. The complete lack of solubility of 3 in THF is important, since the reaction by-products formed in the synthesis of 3 exhibit limited solubility in THF and can therefore be separated from 3 by repeated washings with THF. Recrystallization of [Rh(n3-TMPP)2][BF4]2 also relies on its insolubility in THF; recrystallization is accomplished by concentrating 3:1 CH2012 / THF solutions of 3 on a rotary evaporator until the solutions are sufficiently rich in THF to precipitate the paramagnetic complex as a purple crystalline solid. In contrast to the [BF4]' salt, the [PF6]' salt of [RhH(n3- TMPP)2]2+ is slightly solubile in THF. Remarkably, [Rh(n3-TMPP)2][BF4]2 is air-stable, but CH2012 solutions decompose slowly over periods of several weeks, depositing a finely divided insoluble gray solid presumed to be rhodium metal. Solid samples, on the other hand, are stable for indefinite periods in air. These observations are somewhat surprising considering the documented, sensitive nature of Rh(II) monomers. Typically these molecules can only be synthesized by electrochemical methods and subsequently decompose within short periods of time even under anaerobic conditions.22 The infrared spectrum of [Rh(n 3-TMPP)2][BF4]2 displays bands assignable to coordinated TMPP ligands and a broad feature at 1050 cm-1 which is indicative of the [BF.,]' counterion. The electronic spectrum recorded in CH2012 exhibits a d-d transition at Am” = 537 run (2050 M-1 cm'l) which is responsible for the intense purple color of the compound. Several higher 71 energy charge-transfer transitions are located at Amax = 329 nm (1.34 x 104 M- 1cm-l), 298 nm (1.53 x 104 M-1cm'1) 285 nm (1.75 x 104 M'1cm'1), 233 nm (6.24 x 104 M-1cm-1), 420 nm (sh). Not unexpectedly for a d7 metal complex, the 1H and 31P NMR signals are broad and reveal no information about the structure. In general, Rh(II) phosphine complexes have been characterized by their EPR activity and their broad, largely unidentifiable NMR features. We note that this data and ours are in sharp contrast to several puzzling instances in which paramagnetic Rh(II) compounds have evidently exhibited sharp and unshifted NMR signals as well as EPR signals.3°:23 B. Magnetic and EPR Spectroscopic Properties of [Rh(n 3- TMPP)2] [BF412 (3) The paramagnetism of 3 was investigated by EPR spectroscopy and magnetic susceptibility, the results of which are presented in Table 7. EPR measurements of 3 were carried out in a variety of solvents at 298 K and at 77 K. At room temperature, solutions of 3 in CH2012, acetonitrile/toluene (1:1) and CH2012/Me-THF (1:1) produce identical spectra with a single isotropic homogeneous line at g z 2.20. The EPR spectrum of [Rh(n3- TMPP)2][BF4]2 in a 0H2012/MeTHF glass at 77 K exhibits a rhombic signal as shown in Figure 8. The g—values of the CH2012MeTHF system are gxx = 2.26, gyy = 2.30 and gzz = 1.99. The proximity of gzz to that of the free- electron value is consistent with a dz2 ground state for the unpaired electron and confirms the radical nature of 3. The average of the three anisotropic g- values equals that of the isotropic line at room temperature, therefore the change in line shape is due to the rapid tumbling of the complex in solution, .m E as 33» BE. 525636 d 3 5 «3833329355: .5 8583 Em -: u I..- mmzmo «N l_ 72 73 as opposed to being frozen in random orientations with respect to the applied field at 77 K. In the anisotropic spectrum (77 K), gzz is split into a doublet (Au: 2.0 x 10'3 cm'l) due to the hyperfine interaction with the I=1/2 nucleus of 103Rh. Since the linewidths along gxx and gyy are broader than the hyperfine tensor components along these axes, hyperfine coupling is not observed. Similar behavior has been reported for Rh(II) species trapped in zeolites?4 In the zeolite study, a doublet splitting of gzz of ~32 G was reported and a dynamic J ahn-Teller effect was observed at high temperatures (400°C). Even so, examples of 103Rh hyperfine coupling along all three principal directions are not unknown; Wilkinson and co-workers observed hyperfine coupling along all three axes (Ax = 4.7 x 10'3 cm'l, Ay = A2 = 5.6 x 10'3 cm'l) in the organometallic species Rh(2,4,6-Pri306H2)2(tht)2.25 The magnetic moment of [Rh(n3-TMPP)2][BF412 (3) was measured both in solution and the solid state. The results of these studies are listed in Table 7. Solid state susceptibility measurements were performed over a temperature range of 5-286 K at a variety of field strengths. Compound 3 exhibited Curie-Weiss behavior over the entire range of temperatures. A plot of Xm vs 1/T yielded a temperature independent paramagnetism (T.I.P.) contribution of 424 x 10'6 cgsu, which was applied, along with a correction for diamagnetism, to the average magnetic moment giving a value of 1.80(3) HB- Solution studies were carried out using the Evans NMR method26 and resulted in a Heff of 2.10 03. Both measurements are consistent with a S=1/2 ground state, as is expected for a low spin d7 ion. Similar values have been reported for several other mononuclear Rh(II) complexes? 74 Table 7. EPR Spectroscopic and Magnetic Susceptibility Data for [Rh(n3- TMPP)2][BF412 (3) [Rh(n3-TMPP)2][BF4]2 (3) EPR (CHzClg/Me-THF glass)a gxx 2.26 gyy 2.30 gzz 1.99 Azz, G (cm-1) 22 (2.0 x 10-3) magnetic moment, peg, 1113b solution0 2. 10 solidd 1.80 3 Measured at 77 K. b A diamagnetic correction of -724 x 10'6 was applied based on -20 x 10'6 for Rh2+, -39 x 10'6 for BF4' and -313 x 10'6 for TMPP. 0 Measured in CH2012 at 294 K using Evans methodls. d Average magnetic moment measured over a temperature range of 5-286 K at several field strengths. 75 0. Crystal Structure of [Rh(n3-TMPP)2][BF4]2 (3) The molecular cation shown in Figure 9, consists of two phosphine ligands bonded to the Rh atom in a face capping tridentate mode through the oxygen atoms of two pendant methoxy-groups and the phosphorus atom. This bonding arrangement of TMPP is identical to that observed in the molybdenum tricarbonyl complex (n3—TMPP)Mo(CO)3?7 The geometry about the Rh center is a distorted octahedron, with the Rh atom residing on a two- fold axis that bisects the P-Rh-P' angle (Figure 10). Perhaps the most striking feature of the structure is the fact that the phosphines lie cis to each other rather than trans, as expected on the basis of steric arguments. The consequences of steric repulsion resulting from the cis geometry are evident in the P(1)-Rh-P(1)' bond angle of 105.2(1)°. Indeed Rh(II) complexes with the formula RhX2L2 where L is a bulky tertiary phosphine have been shown to exist in a transconformation.6 While it is conceivable that the cis complex is a kinetic product, we have been unsuccessful in isomerizing the compound to a transgeometry at higher temperatures. It is interesting in the context of this discussion to note that Rauchfuss et al. also observed a cis arrangement of PPh2(o-Me006H4) groups for Ru012(PR3)2, in which the ether-phosphine also participated in a chelating interaction with the metal?3 Furthermore, Shaw noted similar cis isomers in the reaction of PPh2(06H4OH) with Pd and Pt chlorides.29 Apparently, electronic factors play a prominent role in stabilizing the cis conformation relative to the sterically favored trans geometry for these mixed ether-phosphine donors. In the case of 3, the chelation results in the formation of a five- membered metallacyclic ring, the geometric requirements of which distort the molecule as evidenced by the strained P(1)—Rh(1)-O(6) bond angles of 77 .7(2)° and 80.5(1)° (Table 8). This distortion is of the same magnitude as that 76 C(17) 0(5) 0..— C(16) C12 C(9)0 ( ) 0(4) C(26) C(22) Q) C(27) C(23) 0“” (Iowan) 3 50(21) ., t’——-OC(25) C(19) C(4)) 0(7) W Figure 9. ORTEP representation of the [Rh(n3-TMPP)2]2+ (3) molecular cation with 40% probability ellipsoids. Phenyl ring atoms are shown as small spheres of arbitrary size for clarity. 77 Figure 10. ORTEP drawing emphasizing the coordination sphere of [Rh(n3-TMPP)2]2+. 78 observed for other complexes containing similar chelate rings, such as As013(PR3)230 Ru012(PR3)228 and (n3-TMPP)Mo(CO)327 The distances within the immediate coordination sphere of Rh are worth commenting upon as they show some unusual features. The Rh(I)-P bond distance of 2.216(2)A is significantly shorter than the corresponding distances found in most Rh(I) and Rh(III) phosphine complexes, which are generally in the range of 2.28 A to 2.37 A31 As for the Rh-P bond distances in other mononuclear Rh(II) compounds, the only other example which we are aware, trans-Rh012(PPri3)2, exhibits Rh-P bonds that are slightly longer, 2.366(1) AG» 32 This is not unexpected, since in Rh012(PPri3)2 the phosphines exert a significant trans effect on each other. In contrast, complex 3 contains ether donors in positions opposite to the phosphorus donors; these would be expected to exhibit little if any trans effect.33 Indeed the Rh-O bond distances in 3 are quite long as compared to those of metal alkoxides or other anionic oxygen donor ligands and are indicative of the weak nature of the Rh- ether bond.34 Rauchfuss arrived at a similar conclusion upon observing that Ru012(PR3)2, with phosphorus atoms trans to methoxy-groups, exhibits Ru-P bond distances that are much shorter than in other Ru(II) phosphine complexes?8 Of further importance in the solid state structure of [RhH(n3- TMPP)2]2+ is the presence of two chemically distinct metal-ether interactions. The oxygen atoms that are trans to phosphorus are bonded at a distance of 2.201(6) A, whereas the mutually trans oxygen atoms exhibit Rh-O distances of 2.398(1) A. The ain'al elongation is undoubtedly the result of the unpaired electron residing in the dz2 orbital and constitutes the first structural evidence for such an effect in d7 Rh(II) chemistry. Removal of the odd electron via chemical oxidation results in shortening of the Rh-OMe ‘l 79 GKNG 85 85 85 852G 855 GE :55 58.8: 85 85 85 858G 8G 5 G5 G5 ASYMNG 85 85 G5 85883 830 G5 G5 vadNG 85 85 G5 Gums—«NH 850 G5 GEM 88.8: G5 85 G5 88.8 855 GM GEM End: 85 G5 85 85%: G5 G5 GEM 858G 85 G5 GE 85?: .85 GEM 85 8:.NNG 85 G5 GM 858 .85 GEM G5 888$: 8N5 85 8N5 85$ 85 GEM G5 888.8: 8N5 85 8N5 85an .G5 GEM G5 $8.8: 8G5 85 850 8x32 .85 GEM G5 858: 8G5 85 GEM 852. 85 GEM G5 858G 8:0 85 GG5 GER: .G5 GEM G5 888: 85 85 85 G53 G5 GEM G5 588: 85 85 85 G583 .GE GEM GE 3ow m 888 N 83m G 885 2ow m 880 N 838 G 83m 3&5» 83M GM: 85 G5 858G 85 85 GEM: 85 G5 GVEG 85 85 8%.“va 8N5 85 85%: 3.5 85 8:82 8N5 85 85$.G EU G5 G5: 8N5 85 88%: 85 G5 8:82 :55 G5 855G :50 GE GEvG 8:0 85 GENwG 850 G5 GmeG 850 85 GENwG G5 G5 G55 8:5 85 8MoN.N 85 GEM G52 G55 85 fivwmmN G5 GEM 853G 85 85 ANEGNN G5 GEM 005328 N 838 G 888 853me N 838 G 88m magma—ma 98M .5 agmfianzessia as are .555 use 80 36:35:. 9.8. 3823 .m 933... 80 distances, so that both sets of methoxy bonds to the metal, axial and equatorial, are essentially equivalent in length (vide infra). Interestingly, although the axial ether groups are located at long distances, they afford protection from attack by donor solvents such as MeCN or oxidizing solvents such as CH2Clz. Similar protective properties of ortho methoxy substituents were also noted by Rauchfuss?8 D. Redox properties of [Rh(n3-TMPP)2][BF412 (3) A cyclic voltammogram of [Rh(n3-TMPP)2][BF4]2 (3) (shown in Figure 11) exhibits two accessible redox couples, one reversible oxidation at Em = + 0.46 V and a quasi-reversible reduction at E1/2 = ~0.65 V (vs Ag/AgCl). Accordingly, we expected that 3 would undergo chemical redox reactions to form Rh(I) and Rh( III) compounds with minimal structural rearrangement. (l) Oxidation of [Rh(n3-TMPP)2][BF412 (3) (1) Synthesis and spectroscopic characterization of [Rh(n3- TMPP)2][BF4]3 Oxidation of 3 with either [NO][BF4] or [0p2Fe][BF4] in MeCN or 0H2012 solutions results in the formation of a deep red solution of [Rh(n3- TMPP)2][BF4]3 (4). The product was isolated as an oily solid by addition of Et20 or evaporation of the solvent. Complex 4 is both air sensitive and thermally unstable. The latter point is illustrated by the fact that solutions of 4 readily decompose at ambient temperatures (vide infra). To circumvent this problem, the synthesis of 4 was carried out at -40°C. The cyclic voltammogram of 4 is identical to that of 3 except that the two redox couples correspond to one-electron reductions to form Rh(II) and Rh(I) species (Figure 12). Shriieder and Cooper reported analogous electrochemical behavior for the homoleptic Rh(III) thioether complexes [Rh(9S3)2]3+, [Rh(lOS3)2]3+ and [Rh(1283)2]3+, for which both the Rh(II) and Rh(I) 81 Elam) = + 0.46 v RhIII / RhII SpA Rh“ / Rh! E1/2(red) = ' 0.65 V _1 1 +1.5 +1.0 +0.5 do -o.'5 -|:O -u.'5 VOLTS vs Ag/AgCl Figure 11. Cyclic voltammogram of [RhH(n3-TMPP)2][BF412 (3) in 0.1 M TBABF4 in CH2C12. 82 oxidation states are accessible.22""c Interestingly, acetonitrile solutions of 3 may also be oxidized to 4 by the addition of an excess of HBF4 - Et20. The reaction, which occurrs slowly over a period of 12 hours, was monitored by both 1H and 31P NMR spectroscopy. Presumably, H+ is reduced to H2 during the course of the reaction, although the formation of H2 was not readily apparent from the 1H NMR spectrum. Infrared spectral measurements of 4 show a typical pattern for coordinated TMPP in addition to a strong broad band at 1050 cm'1 which corresponds to the [BF4]' counterion. The 1H NMR spectrum of 4 in CD2012 exhibits nine resonances between 5 = 2.93 and 5 = 4.69 corresponding to the 12 methoxy substituents for two magnetically equivalent TMPP ligands. Further downfield there are a set of six poorly resolved doublets of doublets of equal intensity found in the region of the meta-protons of the phenyl rings. The ABX pattern of the resonance corresponds to coupling of one ring proton to the other ring proton as well as to the phosphorus. The 31P NMR spectrum in CD2012 shows a doublet centered at 5 = + 37.4 ppm (JRh-p = 107 Hz), further indicating that the two TMPP ligands are equivalent. Therefore, it appears that the ABX pattern of the meta-protons arises from coupling of proximal and distal protons on the individual chelating rings, which then further couple to a phosphorus nucleus to produce the observed ABX splitting pattern. Although only two rings participate in the chelation to the metal, it is apparent that the meta-protons on the third non-interacting ring are magnetically inequivalent as these protons also exhibit an ABX splitting pattern. These results suggest that the non-interacting ring is sterically locked into an asymmetric environment and is unable to rotate freely about the P-C bond, thereby rendering the ring protons inequivalent. 83 EH2 (red); = +0.46 V EH2 (red)2 = -0.65 V I SpA 1 l 1 +15 +10 +05 do -0.' -If0 -I.'5 VOLTS VS Ag/AgCl Figure 12. Cyclic voltammogram of [Rhm(n3-TMPP)2][BF4]3 (4) in 0.1 M TBABF 4 in CH2C12. 84 (ii) Crystal Structure of [Rh(n3-TMPP)2][PF6]2[BF4] Difficulties encountered while modeling a disorder of the [BF4]- anions have precluded a satisfactory refinement of the structure in these regions. Attempts to model the highly disordered counterions are described in the experimental section. These problems notwithstanding, the Rh(III) molecular cation itself is well behaved, and therefore, a brief description of this species is presented here. The preliminary X-ray results show that the Rh atom lies on a two-fold axis that bisects the cis phosphine ligands as in the structure of the parent complex [Rh(n3-TMPP)2]2+ (3) (Figure 13). As a result of this symmetry, the two phosphines are crystallographically equivalent just as was observed in solution by 1H and 31P NMR studies. Both TMPP ligands bind to the metal in the familiar capping tridentate fashion observed in the structure of 3 and in the molecule (n3-TMPP)M0(CO)3?7 The most striking feature of the structure 4 compared to 3 is the dramatic contraction of the axial Rh-OMe bonds while the equatorial distances remain essentially unchanged (Table 9). The Rh(1)-O(1) bond distance has shortened from 2.398(5) A in [Rh(n3-TMPP)2][BF4]2 to 218(2) A in the structure of 4, and is now is roughly equivalent to the equatorial Rh-O(6) distance (222(1) A). Thus, removal of the odd electron from the dz2 orbital of 3 alleviates the observed axial distortion. The above result is important because it illustrates the TMPP ligand's ability to adjust to and accommodate electronic changes at the metal center. (iii) Decomposition of [Rh(n3-TMPP)2][BF4]3 (4) As mentioned earlier, solutions of [Rh(TMPP)2][BF4]3 are thermally unstable with respect to loss of a methyl group to form [Rh(n3-TMPP)(n3: TMPP-0)][BF4]2 (5) ( n 3-TMPP-O = P{06H2(OMe)3}2{06H2(OMe)2O}) as shown in equation 13. At room temperature, solutions of 4 in CD3CN show 85 0/ Figure 13. ORTEP drawing of the [Rh111(n3-TMPP)2]3+ (4) molecular cation. Phenyl ring atoms are shown as small spheres of arbitrary size for clarity. 85G 85 85 85 GVNG 850 G5 8G5 852 85 85 85 GKG 850 G5 G5 85NG 85 85 G5 G53 850 G5 G5 85: 85 85 G5 8VNGNG 850 G5 GEM 85: G5 85 G5 8563 850 G5 GEM 8MNG 85 G5 85 852: G5 G5 GEM 85G 85 G5 GE 8583 .85 GEM 85 8584 85 G5 GE 8585 .85 GEM G5 85NG 8N5 85 G85 8588 85 GEM G5 85: 8N5 85 8N5 8565 .G5 GEM G5 85: 8G5 85 850 85.8: .85 GEM G5 G5: 850 85 GEM 85.2. 85 GEM GE 85: 850 85 GG5 85.83 .G 5 G EM G5 85G 85 85 85 85.8 G5 GEM GE 88: 85 85 85 85MB .GE GEM GE 2858 m 888 N 833. G 83a 2me m 888 N 838 G 888 % mm_wc< 98M 858G 8N5 85 . 858G 85 85 $84 8N5 85 852 85 85 853G 8N5 85 Gina 85 85 853 8N5 85 853 85 G5 858G 8N5 85 858G 85 G5 852 850 85 85: 850 G5 858G 855 85 85m: 850 GE 88me 850 85 858G G5 G5 853 855 85 GVNNN 85 GEM 858G G50 85 85G.N G5 GEM 85G 85 85 8:88 G5 GEM 8:8...me N 838 G 888 893mm. N 838 G 888 $83.55 wcom .83 588385882935:25 as 33 6.336 was 80 38:36.5 EB 6883 .8 33a. 87 evidence of demethylation within minutes by NMR spectroscopy, but upon cooling to -20 °C the process is slowed considerably Dealkylation eventually occurs regardless of solvent choice, but the reaction is much faster in acetone and acetonitrile than in methylene chloride. The process is quite facile, as evidenced by the fact that even solid samples of 4 slowly convert to 5 over a period of weeks at room temperature. If one heats solid samples of 4 to 50 °C under vacuum, nearly quantitative conversion to 5 occurs within 36 hours. -CH; [Rhm(n3-TMPP)2][BF413 a? [Rhm(n3—TMPP)(n3-TMPP-0)][BF412 <13) 2 2 4 r. t. 5 Compound 5 was prepared in bulk by stirring acetone solutions of Rh(n3-TMPP)2[BF4]3 at room temperature for 12 hours; during this time the solution color changed from deep red to orange. The product was isolated as a pale orange solid by addition of diethyl ether, and while its infrared spectrum is nearly indistinguishable from that of [Rh(n3-TMPP)2][BF4]3 (4), the 1H NMR spectrum reveals the magnetic inequivalence of the two phosphine ligands (Figure 14). The spectrum, shown in Figure 14, displays nine distinct resonances between 6 = 5.84 and 6.96 with an ABX coupling pattern and a set of three overlapping signals centered about 5 = 5.60 ppm. Upfield resonances between 8 = 4.51 and 8 = 2.94 ppm correspond to the ortho and para methoxy substituents on the phenyl rings and number 17 by integration, as compared to 18 in the parent complex [Rh(n3-TMPP)2]3+, thereby confirming the loss of a methyl group from one of the methoxy substituents. The appearance of the 1H NMR spectrum is highly solvent dependent. .3030 E «33:6 dashing—25:35-2 .3 883m 522 m; 28 E .3253 can 9 mm 1 a. . a. a. - Ir-v...h..p»_...._...._...._.. ._. .._...._ ._. . ._..».>|...._.L> >533)» $2 ééie: .i§?\<,29 SE “Egan... m; N: S; " £5,558: in 5.7.; r o n .8. c... m... 0;... m m o m m.m o.“ 8 . — . . . . _ . . .i—lplplrlrll— 8 jig/xiii} n _— — _: — _ — £585 59: NF 3:05 >359: : u 220:. >359: .m £32 I. 89 The conversion of [Rh(n3-TMPP)2][BF4]3 (4) to [Rh(TMPP)(TMPP- 0)][BF4]2 (5) has been monitored by 1H NMR in d6-acetone, d3-acetonitrile and d2-methylene chloride; each of these studies revealed the onset of new resonances for 5, but none that could be assigned to the protons of a methylated by-product were observed. This observation coupled with the solid state transformation of 4 to 5 is consistent with the formation of a volatile species that we believe to be CH3F resulting from attack by [BF4]- on the complex (eq 14).35 Although uncommon, such transformations are not unprecedented; In fact, several examples of activation of [BF4]' by transition OMe OMe IF _I -1 f" Me F- B‘u" F MeO ,0 F MeO ,0 CH8 F T ,‘..P-Rhm —> ,‘,.P-Rhm + (14) Ar / Ar / BF3T Ar' Ar‘ metal complexes have been reported.36 Generally these transformations involve highly charged electrophilic metal centers. Such complexes are so susceptible to activation that researchers are resorting to highly deactivated fluoro-substituted anions such as [(C6F5)4B]' and [(3,5-CF3-CGH3)4B]‘ in order to stabilize these highly electrophilic metal centers.37 It is apparent that in order to further expand the chemistry of [Rh(n3-TMPP)2][BF4]3, similar strategies must be adopted in our laboratories. The 31P NMR spectral data provide information regarding the coordination geometry about the metal center.3 8 31P NMR spectral measurements of 5 (Figure 14) reveal two resonances with an ABX splitting pattern at 8 = + 37.9 ppm for PA (1JRh-p=138.9 Hz, 2Jp.p=13.73 Hz) and 5 = + 90 31.5 ppm for P3 (1JRh_p=140.4 Hz, 2Jp_p=13.73 Hz). The small magnitude of the P-P coupling is indicative of a cis disposition of phosphines.39 Furthermore, the similarity in the two Rh-P coupling constants suggests that the phenoxide interaction is cis to both phosphorus atoms, in an axial (ax) position relative to the plane containing the phosphorus atoms. Presence of a Rh-phenoxide bond in a trans disposition to one of the phosphorus nuclei would be expected to result in a larger difference in the Rh-P coupling constants. The structure of 5, shown below, is rationalized by considering that a methoxy interaction with a metal center weakens the O-CH3 bond thus rendering the carbon susceptible to nucleophilic attack. It follows that the stronger the metal-oxygen interaction, the more electrophilic the methyl group becomes. Methoxy groups exert a weaker trans influence than the Me Me 72+ in ff) OMe M PI_:,,, 11111in Fo’ _ea —{ yOMe (j): OMe ax-[Rhm(n3 -TMPP)(T1 3-TMP1>-0)]2+ phosphorus atoms, resulting in stronger M-O interactions for the mutually trans methoxy groups than for those trans to the phosphines. It is worth noting that dealkylation of ether groups in methoxy-phosphines has been previously observed, including for a dirhodium carboxylate complex prepared in our laboratories, but in contrast to the present study, these were found to occur under more forcing conditions and in the presence of much stronger nucleophiles.4°’41 The instability of [Rh(n3-TMPP)2][BF4]3 (4) may, in part, be 91 explained by the relatively hard nature of the Rh3+ cation compared to that of other platinum group metals in the +1 and +2 oxidation states. Furthermore, from their work with ether-phosphines, Shaw and co-workers observed that the ease of dealkylation increased with increased steric bulk of the ancillary substituents on the phosphine.41b It is reasonable to assume that the considerable bulk of the methoxy substituted phenyl rings of TMPP contributes to facile demethylation of [Rh(n3-TMPP)2][BF4]3 (4). We found that dealkylation of a coordinated methoxy-group significantly alters the electrochemistry of the complex. Electrochemical measurements performed on CH2C12 solutions of 5 show that the molecule undergoes an irreversible oxidation at Em = +1.55 V vs Ag/AgCl that most likely corresponds to a ligand-based process based on the fact that a number of other TMPP species undergo similar processes near this potential. More importantly, 5 exhibits an irreversible reduction at E10,c = -0.80 V vs. Ag/AgCl with a coupled chemical oxidation wave at -0.02 V vs Ag/AgCl. Chemical reduction with cobaltocene yields a deep red solution that exhibits a broad 1H NMR signal characteristic of a paramagnetic compound. The latter species resembles the product obtained by the direct reaction of the Rh(II) complex [RhH(n3-TMPP)2][BF412 (3) with additional TMPP. In fact, comparison of the EPR spectroscopic properties of this newly formed Rh(II) complex with that of the product obtained from nucleophilic attack on [RhII(n3-TMPP)2][BF4]2 (3) reveal that the two compounds are identical. We propose that the two products obtained by both routes are the Rh(II) analog of 5, i.e. the Rh(II) phosphino-phenoxide complex [Rh(TMPP)(TMPP-0)][BF4]. Details of the chemical and structural relationship between 5 and its Rh(II) analog are presented in chapter VII. 92 (2) Reduction of [Rh(n3-TMPP)2][BF4]2 (3) In light of the synthesis of [Rh(n3-TMPP)2][BF4]2 and [Rh(n3- TMPP)2][BF4]3, isolation of a analogous Rh(I) complex ligated solely by TMPP has proven elusive. Chemical reduction of [Rh(n3-TMPP)2][BF4]2 (3) by cobaltocene in either acetone or acetonitrile produces a red-brown solution within several hours at room temperature. No reaction occurs at temperatures less than -15°C. The product was isolated from THF as a brown diamagnetic solid. Samples were invariable contaminated with minor amounts of salts containing [TMPP-CH3]+ and [Cngo]+. The complex series of multiplets and doublets observed for the meta proton resonances in the 1H NMR spectrum of the solid, clearly indicates that the two phosphines are magnetically inequivalent. Moreover, integration of the spectrum reveals that only 17 methoxy resonances are observed, clearly supporting that a dealkylation reaction occurs following reduction of the metal. This is not surprising, in light of the demonstrated presence of [TMPP-CH3]+ in isolated samples. While the complexity of the 1H NMR spectrum precludes further analysis of the coordination geometry, important information may be gleaned from the 31P NMR spectrum. The 31P NMR spectrum of the product in CD3CN exhibits two resonances with an ABX spin system at 8 = -8.9 ppm (dd, 1JRh_p= 131.8 Hz, 2JpA_PB= 392.4Hz) and -13.4 ppm (dd, 1JRh_p= 120.9 Hz, ZJPA-pB= 392.4 Hz) (Figure 15). The magnitude of the PA-PB coupling constant is in agreement with the presence of two magnetically inequivalent TMPP ligands that have a trans relationship to each other. Based on the the NMR spectral data presented, it appears that the reduction of 3 ultimately results in the formation of trans- RhI(n2—TMPP)(n2-TMPP-O) (eq. 15). Apparently, the Rh(I) cation, [Rh(TMPP)2]1+, formed upon initial reduction of 633283 3V catamaoflmv ”2:531:39fitsAmvmmvmammamzedczia mo 5.5268 05 Scam 838.5 25 mo 83.50on £22 m3 .3 0.3.»er Ema hat mfil mfil VH1 mfil ma! «fin 0H1 m1 ml NI Q! m1 VI _—_______._____——_____——___h—_——_____—__—_____—___________—_—_~_p_—______ A3 2: is. 94 3, is unstable with respect to phosphine dissociation. Upon dissociation, the [Rhn(n3-TMPP)2]2+ fl. "[RhI(TMPP)2]1+" (15a) Me Me O Afar TMPP 0Q?" "[RhI(TMPP)2]1+" Me, 0‘1th 0 (15b) [TMPP-CH3]+ Arm‘ffi Ar' 9 Q Me Me free TMPP then acts as a nucleophile resulting in the demethylation of a coordinated phosphine. Further evidence for phosphine dissociation was provided by the appearance of [TMPP-CHzcl]+ impurities in samples prepared in CH2012. 4. Discussion As far as we are aware, [Rh(n3-TMPP)2]2+ represents the only case of a structurally characterized six—coordinate mononuclear Rh(II) complex. The utility of [Rh2(MeCN)10][BF4]4 as a synthon for unusual mononuclear and polynuclear rhodium species is illustrated by the observation that chemistry of TMPP with other dinuclear starting materials, such as Rh2(OAc)4(MeOH)2, yield only partially substituted dinuclear TMPP complexes“, 42 The stability of the (formally) 19 e' species is attributed to the presence of four interacting ether groups which serve to stabilize the complex both electronically and sterically. The presence of these pendent groups serve two purposes: (1) to kinetically stabilize the metal center from 95 dimerization and (2) to provide thermodynamic stability for the +2 oxidation state relative to either Rh(I) or Rh(III) oxidation states. Without the "built- in" coordinative saturation afforded by the tethered "solvent" molecules, one would expect the molecule to undergo the usual disproportionation or dimerization reactions which are commonly observed in divalent Rh chemistry. However, unlike many complexes kinetically stabilized by sterically encumbering ligand sets, the lability of the ether groups opens up coordination sites so that fiirther reactivity may occur. Moreover, the ready availability of the dangling ether groups allows for reversible substitution reactions with a variety of substrates, a situation which is generally not possible with complexes that undergo dissociative loss of a ligand. We have exploited these aspects of the ligand in order to further develop the potentially rich area of mononuclear Rh(II) chemistry, results of which are presented in subsequent chapters. 10. 96 List of References (a) Kemmit, R. D. W.; Russel, D. R.; in Comprehensive Organometallic Chemistry; Wilkinson, G.; Stone, F. G. A.; Abel, E. W., eds.; Pergamon: New York, 1982; vol. 5. (b) Smith, T. D.; Pilbrow, J. R. Coord. Chem. Rev. 198], 39, 295. (c) Stoppioni, P; Dapporto, D.; Sacconi, L. Inorg. Chem. 1978, 17, 718. (d) Klein, H.-F.; Karsch, H. H. Chem. Ber. 1976, 109, 1453. (e) Mu, X. H.; Kadish, K. M. Inorg. Chem. 1989, 28, 3743. Felthouse, T. R. Prog. Inorg. Chem. 1982, 29, 73 and references therein. (a) Billig, E.; Shupack, S.I.; Waters, J.H.; Williams, R.; Gray HE. J. Am. Chem. Soc. 1964, 86, 926.(b) Pneumatikakis, G.; Psaroulis, P. Inorg. Chim. Acta 1980, 46, 97. (c) Pandey, K. K.; Nehete, D. T.; Sharma, R. B. Polyhedron 1990, 9, 2013. (d) Peng, S. M.; Peters, K.; Peters, E. M.; Simon, A. Inorg. Chim. Acta 1985, 101, L35. (a) Wayland, B. B.; Sherry, A. E.; Coffin, V. L. J. Chem. Soc., Chem. Commun. 1989, 662. (b) Wayland, B. B.; Sherry, A. E. J. Am. Chem. Soc. 1989, 111, 5010. (c) Wayland, B. B.; Sherry, A. E.; Poszmik, G.; Bunn, A. G. J. Am. Chem. Soc. 1992, 114, 1673. (d) Wayland, B. B.; Ba, 8.; Sherry, A. E. J. Am. Chem. Soc. 199], 113, 5305. (e) Sherry, A. E.; Wayland, B. B. J. Am. Chem. Soc. 1990, 112, 1259. (f) Bunn, A. G.; Wayland, B. B. J. Am. Chem. Soc. 1992, 114, 6917. (a) Bennett, M. A.; Lonstafl‘, P. A. J. Am. Chem. Soc., 1969,91, 6266. (b) Moers, F. G.; DeJong, J. A. M.; Beaumont, P. M. J. J. Inorg. Nucl. Chem. 1973,35, 1915. (c) Masters, C.; Shaw, B. L. J. Chem. Soc. (A) 1971, 3679. (d) Empsall, H. D.; Heys, P. N.; Shaw, B. L. Transition Met. Chem. 1978, 3, 165. (e) Empsall, H. D.; Hyde, E. M.; Pawson, D.; Shaw, B. L.; J. Chem. Soc., Dalton Trans. 1977, 1292. (a) Harlow, R. L.; Thorn, D. L.; Baker, R. T.; Jones, N. L. Inorg. Chem. 1992, 31, 993. (b) Rappert, T.; Wolf, J .; Schulz, M.; Werner, H. Chem. Ber. 1992,125, 839. Van Gaal, H. L. M.; Verlaak, J. M. J .; Posno, T. Inorg. Chim. Acta. 1977, 23, 43. (a) Zhilyaev, A. N.; Rotov, A. V.; Kuz'menko, I. V.; Baronovskii, I. B. Doklady Akad. Nauk 1988,302, 275. (b) Pneumatikakis, G.; Psaroulis, P. Inorg. Chim. Acta. 1980, 46, 97 . Dunbar, K. R. Comments Inorg. Chem. 1992, 13, 0000. (a) Dunbar, K. R. J. Am. Chem. Soc. 1988, 110, 8247. (b) Dunbar, K. R.; Pence, L. E. manuscript in preparation. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 21. 22. 97 (a) Dunbar, K. R.; Quillevéré, A. Inorg. Chem. submitted for publication. (b) Quillevéré, A. Ph. D. Dissertation, Michigan State University, 1992. Dunbar, K. R.; Pence, L. E. Inorg. Synth. 1992,29, in press. Gray, H. B.; Hendrickson, D. N.; Sohn, Y. S. Inorg. Chem. 197 1, 10, 1559. Morris, D. E.; Dunbar, K. R.; Arlington, C. A.; Doom, S. K.; Pence, L. E.; Woodruff, W. H. submitted to J. Am. Chem. Soc. (a) Bino, A.; Cotton, F. A.; Fanvvick, P. E. Inorg. Chem. 1979, 18, 3558. (b) Cotton, F. A.; Frenz, B. A.; Deganello, G.; Shaver, A. J. Organomet. ' Chem. 1973, 50, 227. Structure Determination Package: Enraf-Nonius, Delft, The Netherlands (1979). TEXSAN-TEXRAY: Structure analysis package, Molecular Structure Corporation (1985). SHELXS-86: Sheldrick, G. M. in Crystallographic Computing 3; Sheldrick, G. M.; Goddard, R. Eds; Oxford University Press, 1984; p. 175. Walker, N.; Stuart, D.Acta. Cryst. 1983,A39, 158. Morris, D. E.; Dunbar, K. R.; Arlington, C. A.; Doorn, S. K.; Pence, L. E.; Woodruff, W. H. submitted to J. Am. Chem. Soc. For other tetranuclear Rh46+ species see (a) Miskowski, V. M.; Gray, H. B. Inorg. Chem. 1987,26, 1108. (b) Mann, K. R.; DiPierro, M. J .; Gill, T. P. J. Am. Chem. Soc. 1980, 102, 3965. (c) Miskowski, V. M.; Sigal, I. S.; Mann, K. R.; Gray, H. B.; Milder, S. J.; Hammond, G. S.; Ryason, P. R. J. Am. Chem. Soc. 1979, 101, 4384-5. (d) Mann, K. R.; Gray, H. B. Adv. Chem. Ser. 1979, 173, 225. (e) Mann, K. R.; Lewis, N. S.; Miskowski, V. M.; Erwin, D. K.; Hammond, G. 8.; Gray, H. B. J. Am. Chem. Soc. 1977, 99, 5525. Pence, L. E. Ph. D. Thesis, Michigan State University, 1992. Examples of coordination and organometallic Rh(II) complexes prepared electrochemically: (a) Cooper, S. R.; Rawle, S. C.; Yagbasan, R.; Watkin, D. J. J. Am. Chem. Soc. 1991,113, 1600. (b) Rawle, S. C.; Yagbasan, R.; Prout, K.; Cooper, S. R. J. Am. Chem. Soc. 1987, 109, 6181. (c) Blake, A. J .; Gould, R. D.; Holder, A. J .; Hyde, T. I.; Schroder, M. J. Chem. Soc., Dalton Trans. 1988, 1861. (d) Anderson, J. E.; Gregory, T. P. Inorg. Chem. 1989, 28, 3905. (e) Scwarz, H. A.; Creutz, 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 98 C. Inorg. Chem. 1983,22, 207. (e) Mulazzani, Q. G.; Emmi, S.; Hoffman, M. Z.; Venturi, M. J. Am. Chem. Soc. 198], 103, 3362. (f) Kblle, U.; Klaui, W. Z. Naturforsch. 199], 46b, 75. (g) Bianchini, C.; Laschi, F.; Ottaviani, M. F.; Peruzzini, M.; Zanello, P.; Zanobini, F. Organometallics 1989,8, 893. (h) Bianchini, C.; Meli, A.; Laschi, F.; Vizza, F.; Zanello, P. Inorg. Chem. 1989,28, 227. (i) Pilloni, G.; Schiavon, G.; Zotti, G.; Zecchin, S. J. Organomet. Chem. 1977 , 134, 305. (i) Fischer, E. 0.; Lindner, H. H. J. Organomet. Chem. 1964, 1, 307. (k) Fischer, E. 0.; Wawersik, H. J. Organomet. Chem. 1966,5, 559. (1) Keller," H. J.; Wawersik, H. J. Organomet. Chem. 1967,8, 185. (m) Dessy, R. E.; King, R. B.; Waldrop, M. J. Am. Chem. Soc. 1966,88, 5112. (n) Dessy, R. E.; Kornmann, R.; Smith, C.; Hayter, R. J. Am. Chem. Soc. 1968,90, 2001. Ogle, C. A.; Masterman, T. C.; Hubbard, J. L. J. Chem. Soc., Chem. Commun. 1990, 1733. Goldfarb, D.; Kevan, L. J. Phys. Chem. 1986,90, 264; 2137; 5787. Hay-Motherwell, R. S.; Koschmieder, S. U.; Wilkinson, G.; Hussain- Bates, B.; Hursthouse, M. B. J. Chem. Soc., Dalton Trans. 1991, 2821. (a) Evans, D. F. J. Chem. Soc. 1959, 2003. (b) Deutsch, J. L.; Poling, S. M. J. Chem. Ed. 1969, 46, 167. Dunbar, K. R.; Haefner, S. C.; Burzynski, D. J. Organometallics 1990, 9,1347. Jeffrey, J. C.; Rauchfuss, T. B. Inorg. Chem. 1979,18, 2658. Empsall, H. D.; Shaw, B. L.; Turtle, B. L. J. Chem. Soc., Dalton Trans. 1976, 1500. Granziani, R.; Bombiere, G.; Volponi, L.; Panattoni, 0.; Clark, R. J. H. J. Chem. Soc. (A) 1969, 1236. Corbridge, D. E. C. The Structural Chemistry of Phosphorus; Elsevier: Amsterdam, 1974, chapter 11. The structure Rh(PPh3)2Cl(s-qui) (s-qui = semi-o-benzoquinone diimine) of has appeared. However, no magnetic evidence was presented to support the assignment of the rhodium oxidation state: Peng, S. M.; Peters, K.; Peters, E. M.; Simon, A. Inorg. Chim. Acta 1985, 101, L35. Cetinkaya, B.; Lapport, M. F.; McLaughlin, G. M.; Turner, K. J. Chem. Soc., Dalton Trans. 1974, 1591. 34. 35. 36. 37. 38. 39. 40. 41 42. 99 For example see: (a) Chebi, D. E.; Fanwick, P. E.; Rothwell, I. P. Polyhedron 1990, 9, 969. (b) Mehrotra, R. C.; Agarwal, S.K.; Singh, Y. P. Coord. Chem. Rev. 1985, 68, 101. (c) Boyar, E. B.; Robinson, S.D. Coord. Chem. Rev. 1983, 50, 109. (d) Graham, D.E.; Lamprecht, G. J .; Potgieter, I. M.; Roudt, A.; Leipoldt, J. G. Trans. Met. Chem. 199], 193. (e) Bryndza, H. E.; Tam, W. Chem. Rev. 1988,88, 1163. Igau, A.; Gladysz, J. A. Polyhedron 1991,10, 1903. (b) Ferandez, J. M.; Gladysz, J. A. Organometallics 1989,8, 207. (a) Gorrel, I. B.; Parkin, G. Inorg. Chem. 1990,29, 2452. (b) Thomas, R. R.; Chebolu, V.; Sen. A. J. Am. Chem. Soc. 1986, 108, 4096. (c) Crabtree, R. H.; Hlatky, G. G.; Holt, E. M. J. Am. Chem. Soc. 1983, 105, 7302. (d) Hitchcock, P. B.; Lappert, M. F.; Taylor, R. G. J. Chem. Soc., Chem. Commun. 1984, 1082. (e) Reedijk, J. Comments Inorg. Chem. 1982, 6, 379 and references therein. For instance see (a) Brookhart, M.; Rix, F. C.; DeSimone, J. M.; Barborak, J. C. J. Am. Chem. Soc. 1992, 114, 5894. (b) Yang, X.; Stem, C. L.; Marks, T. J. Organometallics 1991, 10, 840. (c) Yang, X.; Stem, C. L.; Marks, T. J. J. Am. Chem. Soc. 1991,113, 3623. (d) Horton, A. D.; Orpen, A. G. Organometallics 1991, 10, 3910. (e) Brookhart, M. ; Volpe, A. F. Jr.; DeSimmone, J. M.; Lamanna, W. M. Polym. Prepr. 1991, 32(1), 461. Meek, D. W.; Mazanec, T.J.Acc. Chem. Res. 198], 14, 226. (a) Pregosin, P. S. Phosphorus-31 NMR Spectroscopy in Stereochemical Analysis; Verkade, J. G.; Quin, L. D. ed., VCH, 1987. (b) Pregosin, P. S.; Venanzi, L.M. Chem. Br. 1978, 14, 276. (a) Chen, S. J .; Dunbar, K. R. Inorg. Chem. 199], 30, 2018. (b) Chen, S. J .; Dunbar, K. R. Inorg. Chem. 1990, 29, 588. (a) Kurosawa, H.; Tsuboi, A.; Kawasaki, Y.; Wada, M. Bull. Chem. Soc. Jpn. 1987, 60, 3563. (b) Empsall, H. D.; Heys, P. N.; McDonald, W. S.; Norton, M. 0.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1978, 1119. (c) Empsall, H. D.; Heys, P. N.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1978, 257. (d) Empsall, H. D.; Hyde, E. M.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1975, 1690. (e) Jones, C. E.; Shaw, B. L.; Turtle, B. L.J. Chem. Soc., Dalton Trans. 1974, 992. (f) Empsall, H. D.; Hyde, E. M.; Jones, C. E.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1974, 1980. Dunbar, K. R.; Saharan, V.; Matonic, J. M. manuscript in preparation. CHAPTER IV REVERSIBLE CARBON MONOXIDE CHEMISTRY OF [Rh(n3-TMPP)2] [BF4] 2 100 101 1. Introduction The synthesis and isolation of [Rh(n3—TMPP)2][BF4]2 provided us with a unique opportunity to explore the chemistry of authentic Rh(II) d7 metallo- radicals. Typically, radical species are highly reactive and only observed as transients resulting from homolysis of M-M and M-L bonds.1 Such species have been implicated in a number of important stoichiometric and catalytic transformations.2 In contrast to the well-documented chemistry of early transition metal radicals, radical species of the platinum group are less understood. Recent work with Rh(II) porphyrin complexes by Wayland and co-workers, which revealed the ability of such species to activate C-H bonds and couple CO grOups, underscores the fact that platinum group radical complexes are as important an area of investigation as their early transition metal counterparts.3 Early studies of Rh(II) phosphine complexes revealed the proclivity of these systems to undergo rapid disproportionation reactions.4 For example, reaction of RhClZ(PCy3)2 with CO initially yields a paramagnetic species believed to be the Rh(II) carbonyl complex, RhC12(PCy3)2(CO), which, in solution, rapidly disproportionates to the more stable Rh(I) and Rh(III) counterparts.4e The initial Rh(II)CO adduct is so unstable with respect to disproportionation that it can only be detected in solid state reactions. One method of preventing such a reaction is by introducing a more rigid ligand environment that disfavors ligand redistribution. This concept is elegantly illustrated by the chemistry of Rh(II) porphyrins. The porphyrin macrocyclic framework prevents ligand redistribution from occurring and therefore kinetically stabilizes the Rh(II) oxidation state with respect to disproportionation. As a consequence, reactivity is dominated by ligand association reactions and not electron transfer processes. When bulky 102 substituents were added to the porphyrin macrocycle, dimerization of the d7 metal fragments was disfavored resulting in the isolation of metal centered radicals.3 Similar results may be obtained for Rh(II) phosphine complexes by incorporating chelating groups into the phosphine that deter ligand dissociation. Some work in the past has supported this concept, for example that of Shaw et al. who used this approach to isolate moderately stable Rh(II) and Ir(II) species with the general formula MH(OC6H4P(But)2)2.5:6 In fact, they found that the Ir(II) complexes formed stable adducts with Oz and CO.7 The corresponding small molecule chemistry of the Rh(II) species was not reported. With these fascinating but limited results as a backdrop, we set out to explore the reactivity of [Rh(n3-TMPP)2][BF4]2 by exploiting the weak nature of its Rh-ether bonds to study its potentially rich small molecule substitution chemistry, particularly with CO, in a quest for forming paramagnetic adducts that may exhibit novel reactivity. This chapter investigates the redox and substitution chemistry of [Rh(n3-TMPP)2l[BF4]2 with CO. 2. Experimental A. Synthesis All reactions were carried out under an argon atmosphere by the use of standard Schlenk-line techniques unless otherwise stated. Reactions at pressures greater than 1 atm were performed in a 450 mL stainless steel Parr mini reactor (model 4560) equipped with a magnetic drive stirrer and an automatic temperature control. The starting materials tris(2,4,6- trimethoxyphenyl)phosphine (1) (TMPP) and [Rh(n3-TMPP)2][BF4]2 (3) were prepared as described in Chapters II and III, respectively. The starting 103 material [Rh(cod)Cl]2 was purchased from Strem Chemicals. [Rh(CO)2Cl]2 was synthesized from [Rh(cod)Cl]2 by displacement of cod with CO in THF. The salt [szFe][BF4] was prepared by oxidation of Cnge with hydrofluoroboric acid in the presence of p-benzoquinone.8 The reagents CNBut, NOBF4, and Cngo were purchased from Strem Chemicals and used without further purification. Carbon monoxide was obtained from Matheson Gas Products and was used as received. Labeled 13C0 (99%) was purchased from Aldrich. (1) Reaction of [Rh(n3-TMPP)2][BF4I2 (3) With CO. In a typical reaction, [Rh(n3-TMPP)2][BF4l2 (3) (0.100 g, 0.07 mmol) in 10 mL of CH2C12 was purged with CO gas for 20 min. While maintaining a CO atmosphere, 20 mL of CO-purged diethyl ether was added to the solution which effected the separation of an oily red solid from a yellow solution. The solution was decanted from the solid, the volume of the solution was reduced to 5 mL and diethyl ether (20 mL) was slowly added to precipitate a yellow microcrystalline solid. The yellow product was filtered in air and washed with 2 x 10 mL of diethyl ether and dried in vacuo; yield, 0.038 g (40%). The red oil was dissolved in 5 mL of CH2C12 and layered with 20 mL of diethyl ether. A red-orange precipitate formed which was filtered off, washed with 3 x 5 mL diethyl ether, and dried in vacuo; yield 0.043 g (41%). (2) Reaction of [Rh(n3-TMPP)2][BF4]2 (3) with 1200/1300 (1:1). A quantity of [Rh(n3-TMPP)2][BF4]2 (3) ( 0.050 g, 0.035 mmol) was dissolved in 10 mL of CH2C12 and the solution was degassed by several freeze/pump/thaw cycles on a high vacuum line. An equimolar volume of 12CO and 13CO was delivered to the reaction flask with the use of a Toepler pump. The pressure above the solution was calculated to be in the range of 1-2 atm at room temperature. The reaction vessel was allowed to slowly warm to room 104 temperature, during which time the solution color changed from purple to murky red. After 30 min, a small amount of solution was syringed out and its infrared spectrum was measured; v(CO) (cm'l): 2011, 1985, 1968 in an approximate 1:211 intensity ratio. (3) Preparation of [Rh(nZ-TMPP)(TMPP)CO][BF4] (7) (i) Reduction of [Rh(n3-TMPP)2][BF4]2 in the Presence of CO. A solution of [Rh(n3-TMPP)2][BF4]2 (0.100 g, 0.07 mmol) and a 3-fold excess of szCo (0.042 g, 0.22 mmol) in 5 mL of CH2012 was stirred under a moderate ‘ purge of carbon monoxide. After 20 min, 30 mL of diethyl ether was added and the solution was cooled to -5 °C for 24 h. The yellow product was collected by suction filtration, washed with diethyl ether and recrystallized from CH2C12 (10 mL) by reduction of the volume and slow addition of diethyl ether (0.5 mL). A crop of yellow-orange crystals was collected by filtration in air, washed with diethyl ether, and dried in vacuo; yield: 0.057 g (60%). Anal. Calcd for C55H66F4P2019BRh: C, 51.50; H, 5.19. Found: C, 50.57; H, 5.02. IR (Nujol, CsI): v(CO) 1958 cm'l; (CH2C12)I v(CO) 1970 cm'l. 1H NMR (CD2012) 5 ppm: 3.48 (s, 36H, o-OCH3), 3.80 (s, 18H, p-OCH3), 6.04 (t, 4JP_H = 2.1 Hz, 12H, m-H). 31P NMR (CDZCIZ) 5 ppm: -11.1 (d, lJRh_P = 128 Hz). Electronic absorption spectrum (CHZCIZ) Xmax, nm (e): 415 sh, 345 (7770), 305 sh, 285 sh, 254 (64,400). Cyclic voltammogram (0.1 M TBABF4 / CH2012, vs Ag/AgCl): Elm”) = + 0.50 V. (ii) Reaction of [Rh(cod)Cl]2 with TMPP in the Presence of CO. A mixture of [Rh(cod)Cl]2 (0.500 g, 1.01 mmol) and AgBF4 (0.395 g, 2.03 mmol) was dissolved in 5 mL THF for 15 min to give a yellow solution. The yellow solution was then filtered through a Celite plug into a 3-neck flask equipped with a gas inlet and an addition funnel containing a solution of TMPP (2.160 g, 4.056 mmol in 30 mL of THF). The entire apparatus was then cooled to 0°C 105 with an ice bath. The TMPP solution was added dropwise over a period of 10 min during which time the solution was gently purged with CO. After addition of the phosphine, the CO purge was discontinued and the reaction was stirred at 0°C under a CO atmosphere for 1 h. The volume was reduced to 15—20 mL and the reaction solution was stirred for another hour at 0°C to give a yellow precipitate. The yellow solid was filtered in air and washed with diethyl ether (4 x 10 mL). The product was recrystallized by redissolving the solid in 50 mL of CH2C12 followed by filtration and addition of 20 mL of THF. The volume of the solution was reduced to 10-15 mL and the solution was refrigerated at -5°C. A large crop of yellow crystalline solid was collected by filtration in air, washed with 4 x 10 mL of diethyl ether and dried under reduced pressure; yield: 2.250 g (86%). (iii) Reaction of [Rh(CO)2Cl]2 with TMPP A solution flask was charged with [Rh(CO)2Cllz (0.100 g, 0.257 mmol), TMPP (0.548 g, 1.03 mmol) and NaBF4 (0.056 g, 0.514 mmol) and 10 mL of MeCN. The cloudy yellow solution was stirred under reduced pressure for 30 min, after which time, the solvent was removed under vacuum. The yellow product was redissolved in 10 mL of CH2012 and filtered through a Celite plug. THF (10 mL) was added to the solution, and the volume of the solution was reduced to approximately 3-5 mL, yielding a yellow crystalline solid. EtZO (10 mL) was added to precipitate additional product. The solid was filtered in air, washed with 3 x 5 mL of EtzO and dried in vacuo; yield: 0.530 g (80%). (4) Preparation of [Rh(TMPP)2(CO)2][BF4] (6) A quantity of [Rh(n2-TMPP)(TMPP)CO][BF4] (7) (0.100 g, 0.08 mmol) in 5 mL of CHzclg was treated with CO gas for 5 min, after which time 30 mL of CO-saturated diethyl ether was slowly added to induce precipitation of the product. The yellow crystalline solid was collected by suction filtration in air, 106 washed with 2 x 5 mL of diethyl ether and dried. Prolonged drying under vacuum must be especially avoided to prevent potential loss of CO. [Rh(TMPP)2(CO)2][BF4] was isolated as its CHgClz solvate; yield: 0.10 g (90%). Anal. Calc'd for RhC12P2057OzoBF4H68: C, 49.05; H, 4.91. Found: C, 48.22; H, 4.96. IR (Nujol, CsI): v(CO) 2006 cm'l; (CHZCIZ): v(CO) 2011 cm'l. 1H NMR (CD2C12) 5 ppm: 3.41 (s, 36H, o-OCH3), 3.79 (s, 18H, p-OCH3), 6.02 (t, 4JP-H = 2.1 Hz, 12H, m-H). 31P NMR (CD2C12) 6 ppm: -23.8 (d, lJRh-P = 116 Hz). Electronic absorption spectrum (CHZCIZ) kmax, nm (e): 438 (3400), sh, 350 (6400), 288 sh, 256 (61,500). Cyclic voltammogram (0.1 M TBABF4 / CH2C12, vs Ag/AgCl): Ep’a = + 0.80 v. (5) Solid State Reactions of (3) - (7) with CO. In a typical experiment, a small amount of finely divided starting compound (5-10 mg) was loaded in a polyethylene vial and placed in a Parr reactor. After several fillings and subsequent purgings with CO, the reactor was pressurized to approximately 50 psi. Reaction times were varied from 15 min to 5 days. After the vessel was depressurized, a small drop of Nujol oil was added and the sample was quickly transferred to C51 plates for infrared spectral measurements. Alternatively, the reactions were performed by purging a finely divided sample suspended in Nujol with CO at atmospheric pressure for an extended period of time. (6) Redox Reaction of [Rh(n2-TMPP)(TMPP)CO][BF4] (7 ) with [Rh(n3- TMPP)2l [BF4ls (4). A equimolar mixture of [Rh(TMPP)zCO][BF4] 7 (0.012 g, 0.009 mmol) and [Rh(n3-TMPP)2][BF4]3 4 (0.013 g, 0.009 mmol) was dissolved in 10 mL of CH2C12 and the reaction was stirred for 90 min under an argon purge to liberate CO. During this time, the solution color changed from red to red- purple. After stirring overnight under an Ar atmosphere, the solvent was 107 evaporated and a crop of purple crystals was collected, washed with a mixture of CH2C12/Et20 (1:1 v/v) and dried in air. The product was identified as [Rh(n3-TMPP)2][BF4]2 by a comparison of its electrochemical and infrared spectral properties to those of an authentic sample; yield: 0.011 g (90% based on [Rh(nZ-TMPP)(TMPP)CO][BF4l). (7 ) Preparation of [Rh(TMPP)2] [BF4] (3) from [Rh(TMPP)2CO][BF4] (7) A solution of [Rh(TMPP)2(CO)][BF4] (0.500 g, 0.390 mmol) and [Cp2Fe][BF4] (0.107 g, 0.390 mmol) in 50 mL of CH2C12 was purged with N2 overnight at -15°C. CH2C12 was added periodically to maintain the reaction volume at approximately 50 mL. During this time, the solution color slowly changed from red to purple. After the reaction was complete, the solvent was evaporated and the crude purple solid was recrystallized as previously described in Chapter III; yield: 0.460 g (88%). (8) Reaction of [Rh(TMPP)2CO][BF4] (7 ) with small molecules (i) N2, 02, C02, H2 . In a typical experiment, a solution of [Rh(TMPP)2(CO)][BF4l (7) (0.025 g, 0.020 mmol) in 5-10 mL of CH2C12 was gently purged with the appropriate gas at r.t. for 15-30 min. An aliquot was removed by either syringe or cannula and transferred to a solution IR cell. For each of the above gases, only one carbonyl stretch (v(CO) = 1970 cm'l) was observed indicating the presence of unreacted [Rh(n3-TMPP)2][BF4]2 (7). (ii) CNR: preparation of [Rh(TMPP)2(CO)(CNBut)][BF4] (8). To a 5 mL solution of [Rh(TMPP)2(CO)][BF4] (0.250 g, 0.20 mmol) in CH2C12, was added 22 ML of CNBut. The yellow solution was stirred for 15 min. at r.t. A yellow crystalline solid was precipitated by slow addition of 40 mL of diethyl ether. The solid was filtered in air, washed with 3 x 5 mL EtZO and dried in vacuo; yield: 0.251 mg (94%). IR (CH2C12): v(CO) 1998 cm‘l, v(CN) 2188 cm'l. 1H 108 NMR (CD2C12) 5 ppm: 0.96 (5, -But), 3.36 broad (s, -OCH3), 3.78 broad (s, -OCH3), 5.99 broad (s, m-H). 31P NMR (CD2C12) 8 ppm: -23.8 (d, lJRh-P = 116 Hz). Cyclic voltammogram (0.1 M TBABF4 / CH2C12, vs Ag/AgCl): E1/2(ox) = + 0.47 V, Ep,c = -o.01 V. (iii) pyridine. An amount of [Rh(TMPP)2(CO)][BF4] (50 mg, 0.04 mmol) was dissolved in 5 mL of CHZCIZ. One equivalent of pyridine (3.2 uL, 0.04 mmol) was syringed into the solution. The yellow solution was stirred for 15 min at r.t., after which time an infrared spectrum of the solution revealed a single CO band at 1970 cm'1 corresponding to unreacted [Rh(TMPP)2(CO)][BF4] (7). B. X-ray Crystallography The structures of [Rh(TMPP)2(CO)2][BF4 ] (6) and [Rh(TMPP)2(CO)][BF4] (7) were determined by application of general procedures that have been fully described elsewhere.19 Geometric and intensity data for 6 was collected on a Nicolet P3/F diffractometer with graphite monochromated MoKa (la = 0.71073 A) radiation and were corrected for Lorentz and polarization effects. Calculations for 6 were performed on a VAXSTATION 2000 computer by using programs from the Enraf-Nonius Structure Determination Package (SDP) programs.9 Crystallographic data for the compound [Rh(nz- TMPP)(TMPP)CO][BF4] (7) were collected by Molecular Structure Corporation on a Rigaku AFC5R diffractometer with monochromated CuKa radiation and a 12KW rotating anode generator. Calculations were performed by using the Texsan crystallographic software package of Molecular Structure Corporation.10 Important crystallographic parameters pertaining to data collection and structure refinement for compounds 6 and 7 are summarized in Table 10. 109 (1) [Rh(TMPP)2(CO)2] [BF4]-CH2C12 (6). (i) Data Collection and Reduction. Single crystals of 6 that were suitable for X-ray diffraction studies were grown by slow diffusion of Et20 into a CH2C12 solution of 6 under a carbon monoxide atmosphere to prevent loss of CO and formation of 7. A yellow platelet of approximate dimensions 0.89 x 1.0 x 0.37 mm3 was taken up in viscous oil at the end of a glass fiber and placed in a cold stream of N2(g) at -96i3°C. The cell parameters were refined from a fit of 15 reflections in the range 20 .<_ 20 S 30° and indicated a triclinic crystal system; axial photography confirmed the low symmetry of the lattice. A total of 8801 unique data were collected in the range 4 S 20 3 45° by using the co-scan method. No significant decay in the data was noted, as evidenced by three check reflections that were monitored periodically throughout data collection. (ii) Structure Solution and Refinement. The position of the Rh atom was determined from a Patterson Fourier synthesis. The remaining non- hydrogen atoms were located and refined by a series of alternating least- squares cycles and difference Fourier maps. Of the 8801 unique data, 5934 with F02 > 36(F0)2 were used in the refinement of 799 parameters to give residuals ofR = 0.059 and RW = 0.084. (2) [Rh(nz-TMPP)(TMPP)CO] [BF4]°2C¢;H6 (7). (i) Data Collection and Reduction. Yellow platelets of 7 were grown by slow evaporation of a mixture of CH2C12 and CsHs. A small crystal of approximate dimensions 0.25 x 0.10 x 0.10 mm3 was selected and mounted on the end of a glass fiber with epoxy cement. Least-squares refinement of 25 well-centered reflections in the range 58.5 S 20 S 77.20 gave cell parameters that belong to a triclinic cell. Based on an analysis of intensities, 110 Table 10. Summary of crystallographic data for [Rh(TMPP)2(CO)2][BF4] - CH2C12 (6) and [Rh(TMPP)2(CO)][BF4] . 2CeH5 (7). 6 7 Formula RhC12P2F4020C57BH68 RhP2F4019067BH73 Formula weight 1395.73 1432.02 Space group P—1 P-1 a, A 13.318(4) 14.898(5) b, A 13.404(2) 18.060(8) c, A 18.164(4) 14.343(4) (1, deg 95.908(3) 9656(4) [3, deg 97.037(3) 113.84(2) % deg 90.711(3) 104.80(4) V, A3 3200(2) 3308(2) Z 2 2 dcalc, g/cm3 1.448 1.438 M, cm'1 4.74 32.60 Temperature, 0C -96 i 3 23 i 1 Trans. factors, max., 1.00, 0.82 1.00, 0.95 min. Ra 0.059 0.067 wa 0.084 0.069 quality-of-fitc 2.59 2.32 aR=X l lFol' chlIElFol bRw = [2w( | F0 I- We |)2/2w lFo l211/2; w = 1/02( lFo I) cquality-of—fit = [2W( ”:0 I ' l I:c |)2/(Nobs‘Nparameter)lV2 111 the space group was determined to be P-l. Data were collected at 23i1°C by using the co-20 scan technique to a maximum 20 value of 120.3°. Weak reflections, those with F02 < 106(F02), were rescanned at a maximum of two rescans and the counts were accumulated to assure good counting statistics. A total of 10304 reflections were collected, of which 9855 were unique. Intensity measurements of three standard reflections every 150 data points indicated that the crystal had not decayed. An empirical absorption correction based upon azimuthal scans of several reflections was applied to the data; transmission factors ranged from 1.00 to 0.95. Data were also corrected for Lorentz and polarization effects. (ii) Structure Solution and Refinement. The structure was solved by MITHRIL and DIRDIF structure solution programs11 and refined by full- matrix least squares refinement. With the exception of the carbon atoms of the benzene molecule in the lattice, all non-hydrogen atoms were refined with anisotropic thermal parameters. A total of 6393 observations with F02>36(F02) were used to fit 757 parameters to give R = 0.067 and RW = 0.069. The quality-of-fit index is 2.59 and the peak of highest electron density in the final difference map is 1.09 e'/A3. 3. Results A. Reactivity of [Rh(n3-TMPP)2] [BF4]2 (3) with Carbon Monoxide. Solutions of 3 in CH2C12 react smoothly with CO at ambient temperatures and pressures. Within 15—20 minutes, the initial deep purple color of 3 converts to a red-orange hue with concomitant growth of a band at v(CO) = 2011 cm'l. Reactions of 3 with 1:1 mixtures of 12C0 and 1300 gave a three band intensity pattern in the v(CO) region, leading to the assignment of the 2011 cm:1 band to a dicarbonyl species. Subsequent purging with Ar or 112 N2 produces a new species that exhibits a lower energy CO stretch at 1970 cm'l. IR monitoring studies carried out during the period of inert gas purging revealed a direct correlation between the disappearance of the species at 2011 cm'1 and the appearance of 1970 cm'l, without the observation of other CO- containing intermediates. With long purging times, both bands gradually decrease in intensity and finally disappear; at the same time the solution color converts from red-orange to purple. The purple product, obtained as a residue, was identified as the parent Rh(II) complex by cyclic voltammetry, electronic spectroscopy and epr spectroscopy. Intrigued by the indication that we were observing reversible binding of CO to the radical d7 metal complex, we undertook a variety of experiments to elucidate the reaction pathway. B. NMR Spectroscopic Studies. NMR spectroscopy of CO-saturated CD2C12 solutions of 3 revealed that the paramagnetic starting material reacts rapidly to form two diamagnetic species at the same rate. Upon work-up with Et20, bulk reactions in CH2012 produce a red solid and a yellow filtrate. The red solid was identified as [Rh(n3-TMPP)2][BF4]3 4 by 1H and 31P NMR spectroscopies (vide supra). Concentration of the yellow solution produced a crop of air-stable yellow microcrystals displaying an infrared band at v(CO) = 1970 cm—1. The 1H NMR spectrum of 7 in CD2012 shows 3 resonances; a virtual triplet at 5 = 6.04 (4Jp_H = 2.1 Hz) and singlets at 5 = 3.80 and 5 = 3.48 that correspond to the meta protons of the phenyl rings, the p-methoxy protons and the o-methoxy protons respectively. A companion 31P NMR study confirmed that the phosphorus nuclei are equivalent, with a doublet appearing at 5 = -11.1 ppm (J Rh-p = 128 Hz). In light of the solid state structure (vide infra), the 1H NMR can be interpreted to mean that a low energy fluxional process occurs in solution to exchange all ortho-methoxy groups. The exchange mechanism is 113 envisioned to occur through association of a free methoxy group, leading to formation of a 5-coordinate trigonal bipyramidal structure, which is followed by labilization of a methoxy interaction as shown in the schematic diagram below.12 Upon methoxy group dissociation, the phenyl ring is free to rotate about the P-C bond thereby bringing a second o-methoxy group into a proximal position to the metal. Rotation about the Rh-P bond eventually allows all o-methoxy groups on the phosphine ligands to interact with the metal center, thus rendering the PR3 ligands equivalent on the NMR time scale. Variable temperature 1H and 31P NMR experiments reveal that this process continues below -90°C, which attests to the unusually high lability of the ether interactions in Rh-TMPP complexes. Similar behavior has been reported for other ether-phosphine complexes of both early and late transition metals, but unlike the present case, low temperature limiting spectra were observed for these fluxional complexes.13 0 O , I ‘0’21‘35 ‘Q’Ap’\ O’\ Q0 , l ‘\\O‘ I \\O~ q; Pi O R __i_ OC-Rh‘ —¥ I v ‘— v — ‘— Rh OC 1'3 o— 3 0 I ‘0’ Q ko 4v The ease of Rh-O bond dissociation in these complexes is further evidenced by the reaction of [Rh(n 2—TMPP)(TMPP)CO]1+ with a second equivalent of CO. Upon exposure to an atmosphere of CO, a pale yellow solution of [Rh(nZ-TMPP)(TMPP)CO]1+ converts to an intense yellow color, signifying the formation of the trans dicarbonyl species [Rh(TMPP)2(CO)2]1+. This reaction is entirely reversible with loss of carbonyl ligand occurring after m 114 a brief period of purging with Ar or N2. The reversible addition of CO was followed by 1H, 31P NMR, cyclic voltammetry and electronic spectroscopy. 31P{1H} NMR measurements reveal that the 31P resonance undergoes a shift from 5 = -11.1 (JRh-P = 128 Hz) ppm for 7 to 5 = -23.8 (d, JRh_p = 116 Hz) for 6 upon addition of the second carbonyl ligand. The 1H NMR spectrum of [Rh(TMPP)2(CO)2][BF4] (7) in CD2C12 shows magnetically equivalent TMPP ligands with a triplet appearing at 5 = 6.02 (Jp_H = 2.1 Hz) due to the meta protons and two singlets at 5 = 3.41 and 3.79 which integrate in the correct ratio for ortho and para methoxy groups. The symmetrical nature of the resonances indicates that rotation about the Rh-P bond is not sterically hindered by the presence of the two carbonyl ligands. C. Electrochemistry of [Rh(nZ-TMPP)(TMPP)CO][BF4] (7 ) and [Rh(TMPP)2(CO)2][BF4] (6). The cyclic voltammogram of [Rh(nZ-TMPP)(TMPP)(CO)]+ shows a quasi-reversible oxidation at Elam) = +0.50 V (Figure 16c). Not surprisingly, this oxidation process is less accessible than the Rh(II)/Rh(I) couple of the parent complex, [Rh(n3-TMPP)2][BF4]2 (3).14 The addition of a second CO ligand to give the dicarbonyl [Rh(TMPP)2(CO)2]+ results in a large positive shift in the Rh(I)/Rh(II) couple to +0.80 V (Figure 16b) indicative of further destabilization of the Rh(II) oxidation state. The ramifications of this shift on the overall reaction pathway will be detailed in the discussion section. D. Crystal Structures of [Rh(TMPP)2(CO)nl[BF4l (n=1,2) Crystallographic parameters and information regarding data collection and refinement for the structures of [Rh(TMPP)2(CO)2][BF4] (6) and [Rh(TMPP)2(CO)][BF4] (7) are summarized in Table 10. Tables 11 and 12 contain a listing of pertinent bond distances and angles for 6 and 7. at (O) H0 +05 0.0 ~05 -I:o VOLTS VS Ag/AgCl Figure 16. Cyclic voltammograms in 0.1 M TBABF4/CH2C12 for (a) [RhIII(n3-TMPP)2][BF413 (4), (b) [Rh(TMPP)2(CO)2][BF4l (6). (c) [Rh(TMPP)2(CO)][BF4] (7). 116 (l) [Rh(n2-TMPP) (TMPP)CO][BF4l (7). The solid state structure of 7, shown in Figure 17 , consists of a [Rh(nz- TMPP)(TMPP)CO]+ cation and a [BF4]- anion in the asymmetric unit. The geometry about the Rh center is that of a highly distorted square planar arrangement of ligands with trans phosphine ligands and a CO ligand that is approximately trans to an oxygen atom from an interacting o-methoxy group (O(1)-Rh(1)-C(1)=150.2(4)°). To effect this bonding, a highly strained five- membered chelate ring is formed: M-P-C-C-O (P(1)-Rh(1)-O(1)=78.1(2)°). The presence of this rather acute angle results in a gross distortion of the structure from an ideal square planar arrangement. Related molecules with this ligand and other ether-phosphines show similar structural features.15 The distance for Rh(1)—O(1) in the present case of 2.319(7) A is intermediate between the values for the axial and equatorial Rh-O distances found in [Rh(n3-TMPP)2][BF4]2 (3). The observed high lability of the ether group in solution reflects the weakness of the bond. Of further interest in the crystal structure of 7 is the presence of a second methoxy group from the monodentate phosphine ligand at a distance of 2.611(7) A, which is outside the sum of the covalent bonding radii for Rh and 0. As the PLUTO drawing in Figure 18 emphasizes, this oxygen atom is poised to occupy an equatorial site of a trigonal bipyramidal structure. We rationalize that, in solution, the methoxy group readily coordinates to the metal to form such a five coordinate intermediate in order to exchange all six ortho-methoxy groups in a fluxional process (vide supra). (2) [Rh(TMPP)2(C0)2l[BF4l (6)- Compound 6 crystallizes as a symmetrical molecule ligated by two phosphine ligands and two carbonyl groups in a square planar geometry. As 117 C(17) «fig/C C(18) pom) 0(5) C(15) 0(2) C10 duh) C(21) C(22) fl\ 0(8) 0(4) )1.— jP—C(19)C(24)fi(23‘)j\j C(26) C(8) 0(3) 1) 0(9) C(27) “5% C(19) ch C(18) £33066) C(45) C(32)C(33) / C(12) Gamma“) 4(7 P(2) C(15) C(42) C(41) C(11) ’ M C(44) cl M C(30) C(29) W , C(14) C(10) C(50) 0““) C(39) C(34) C(49) ~ 0(16) C(53) Figure 17. ORTEP representation of [Rh(TMPP)(TI2-TMPP)(CO)]1+ (7 ). 118 Figure 18. PLUTO drawing of [Rh(TMPP)(nZ-TMPPXCOH 1+ (7) emphasizing the coordination geometry about the Rh atom. “§ 85.84 :30 3.6 ANVO god: 830 ENE GEM G5: $30 30 $0 85m: $80 ENE .GEM $3.: Ago 85 8.5 GEN: $60 ENE GEM 95de 8va N5 GVO GENE 850 GM G5 53: 890 50 G5 $3: 830 GE G5 53.: 890 65 G5 $53G 8:0 GE GE GKMNG 8N5 GVO ANVO G36: 350 GM GEM $55: 3480 GE GE 8E5: 850 GE GEM $3de GNVO GVO GE 853G Gvo GE GEM $5.me 880 $5 830 3&de 630 GEM G5 $54.“: $50 85 85 Amvmdw $va GEM ANE GESNG ECO ANVO G80 ANVNNS GVO GEM ENE :23: GCO G5 a5 8E8 330 GEM GE SEGNG 35 G5 GEM NEE. GVO GEM GE Gamma 85 G5 GEM GEES ENE GEM GE m 298 m. 83m N 83m G 88w Ewan m 88m N 835 G 83m 1 3%: 98M GKmG G45 G30 vawG €va ANVM as: 85 A80 GENE 880 am G5: 85 G5 avawG 850 GE GE»: ANVO Gvo GMwG 830 GM GEGG GEO 5C0 GMwG GVO GE GEE $5 85 5:3 6:0 GEM GE»: 85 85 633 G5 GEM GEL 85 85 am: 690 GEM ANVme GVVO ANVO AmvvmmN ENE GEM G5: EU GVO QUEEN GE GEM 8:326 N 838 G 88c 893va N 83m G Scam .5 £38 . EmfiooEmmzeEE re 33 East es... 9: 803:6 EB escrow 8553me 98M .2 3.3. 120 was observed in the structure of 7, the two TMPP ligands are trans to each other, but in this case each TMPP is bound through only the phosphorus atom (Figure 19). Unlike for the mono-carbonyl compound, the closest approach of ortho-methoxy groups is well outside the coordination sphere of rhodium, and there are no major structural distortions as evidenced by the nearly ideal angles P(1)-Rh(1)-P(2) = 178.46(6) A and C(55)-Rh(1)-C(56) = 179.0(3) A. The second CO ties up the fourth coordination site thus eliminating the need for additional donation from an ether oxygen atom. Both the Rh-P and Rh-C bond distances fall well within the range of values expected for Rh(I) complexes and are in themselves quite unremarkable. E. Solid State Reactions of (3) - (7 ) with Carbon Monoxide. Finely divided samples of [Rh(TMPP)2(CO)][BF4] susPended in mineral oil react with CO after purging for several minutes as evidenced by the appearance of a second CO band corresponding to [Rh(TMPP)2(CO)2][BF4]. The process, however, is slow and not quantitative as a result of poor CO diffusion into the sample. The transformation can also be effected by pressurizing powder samples with CO at 50 psi. The percent conversion of 7 to 6 was improved by exposing the sample for longer periods of time. Presumably, this is due to increased amounts of CO diffused into the solid. Just as was observed in solution, replacement of the CO atmosphere by an inert gas results in loss of coordinated CO and reformation of 7. Solid samples of [Rh(n3-TMPP)2][BF4]2 (3) also react with CO in much the same manner as was observed for solutions of 3. Samples of 3 pressurized to 50 psi for extended time periods partially convert to [Rh(n2- TMPP)(TMPP)(CO)][BF4] and [Rh(TMPP)2(CO)2][BF4] as evidenced by the appearance of weak CO bands at 1958 and 2006 cm‘1 in the IR spectrum. 121 C(27) C(9) ‘ 1 C(23) C(8) as . C(5 ’. 0(3) 4' C(22) ' 0‘2) C(17) 0(9) (3' C(26) 0(5) C(24) » o. C(13) C(14) C(4) C(8) r C(15)C(18)C(19) 0(3) \ ,- .‘ C(12) '2 N1. "‘05) ' C y — C(11)C(10“')P)(1 C(20) (16) 0(4) '. ‘ 0(1) 0(7) I 0(7) ? *" Rh(1) 19) / ORG/‘1», C(25) ,‘ C(36) 1“”? C(55) C(45)! O(12) C(55) r) 0(15) C(33) C(16) C(52) C(32) .C(28) P9 ) 4 C(41) " k ’ 0‘18’9’7‘14 "‘ C(49) “ ‘h t 'i‘ " 5 C(35) C(29)" C(51) 4- 1: SI . 54 C(11) ,-‘ ) C(50) O(17)"C,5) C(40) \‘ ’) C(30) C(10) ‘ 014 ‘ - C(38) 3. ( '\ n C(13) C(34) C(39) C(43) ;, g C(44) Figure 19. ORTEP drawing of [Rh(TMPP)2(CO)2]1+ (6) with 30% probability ellipsoids. 122 SNVO $va 3:..qu GEM ANNE: €va ANE GEM EVNNG $50 $30 GEM ANVNNG 6N5 ANVM GEM ANvmdoG G30 ANE 8N5 ANENS 850 GE SUD ANEGOG €90 ENE 8N5 ANvmdoG 8C0 GE GVO ANVoGoG 830 NE GEM ANVNdoG 850 GE GVO $56G 85 $5 $0 $de 850 GM GEM GEMNG A30 A30 85 N52: 850 GE GEM GEMS $5 $0 ANVO 8562 G 5 GE G EM GENNG 85 go ANVO 8565 $80 GEM 390 $3: $0 85 ANVO ANvmdw . 880 G EM ENE 3V mm: 85 G5 ANVO Awadw $80 GEM ENE QEGNG 85 G5 GE gnaw $30 GEM GE GVNNNG ANVO Gvo GE 55.8 330 GEM GE ANvodG 8va NE 980 $334.qu ANE GEM GE Ewan m 88m N 83m G 83m Emma m 88m N 83m G 83m mafia—«4 98M vame $0 85 @mmmG ANVU G5 €53; GUO ANVO vaNwG 8&0 ENE AtmmmG G5 G5 fivawG 380 NE EmovG ANVO GVU @meG 8N5 ENE SEEM 8va 8N5 8%”va $50 GE vamNGG 830 8:0 353G 850 G5 8%.va 85 85 AmvawG GVO GE $3me $5 85 £383 880 GEM vavaG A80 85 8vaG 3&0 GEM 8%.va 3.5 ANVO GVNmmN ENE GEM GER: A50 GVO GKNmN GE GEM magmas N 88w G 88m 8535mm. N 83m G 83a 3083me 98m .Ae «6&5 . EEGNSQEEEEE he Amos 8&5 28 2V 883% BB 882% .NG 035,—. 123 Interestingly, upon exposure of 3 to CO for longer than 16 hours, a very weak high energy CO stretch was observed at 2088 cm‘l. This species is directly related to reaction of CO with [Rh(n3-TMPP)2][BF4]2 , as neither samples of [Rh(TMPP)2(CO)][BF4] nor [Rh(n 3-TMPP)2][BF4]3 produced this band under identical conditions. 4. Discussion The reversible CO chemistry of [Rh(n3-TMPP)2][BF4]2 at ambient temperatures and pressure proceeds through a series of reactions that involve the formation of Rh(III) and Rh(I) complexes. The interrelationships of the five compounds involved in the cyclic reaction are depicted in Figure 20. These conclusions were arrived at by establishing independent synthetic routes to intermediates 4 - 7 and subsequently investigating their spontaneous redox behavior and reactivity with CO. Figure 21 outlines rational pathways to the key compounds 4 - 7 as well as their interconversions and reactions with CO. Curiously, although [Rh(n3-TMPP)2]2+ undergoes ether-group dissociation and subsequent cis to trans isomerization in favor of n-acceptors such as CO and CNR (R = Me, Pr-i, Bu—t),44 it is unreactive towards donors such as 02. This is evidently a consequence of the electronic environment of the metal rather than a steric effect, since identical behavior has been noted for ether-phosphine complexes in which the ligand is not particularly bulky.15b There are several key features of the chemistry in Figure 20 that must be emphasized in any discussion of this work. The initial attack of CO on the Rh(II) monomer 3 is slow, therefore unreacted starting material is in large excess during the early stages of the reaction. While we never actually in 124 _ OI _ A 2+ 2+ (50 :3 2 f 9,! (loch?) -L->O E: ”I'll 8 [Rhnm3 -TMPP)2]2+ (3) —3[Rhn(n -'I‘MPP)2]2+ (3a) [Rhfl(“3‘TMP ”212+ ‘ -Co [RhH(113-TMPP)2]2+ (3) (3) [Rhm (113-Twang]3+ [Rhmm3_TMPP)2]3+ <4) (4) E1/2(red) = +0.46 V E1/2(ox) = +0.46 V \ 1+ ‘0 1 1+ 0 0' —| I f 8 O [9,“!th O) 0' E €914.1th J 0.. O V I o | ' P\/ 7 ' I 5‘) I IO +CO 8Q [RhI(TMPP)2(CO)]1+ [RhI(TMPP)2(CO)2]1+ (7) (6) v(CO) = 1970 cm‘1 v(CO) = 2011 cm'1 E1/2 (ox) = +0-50 V 131/2(0):) = +0.80 v Figure 20. Proposed pathway for the reversible reaction between [RhH(113-TMPP)2]2+ (3) and CO. Figure 21. High yield synthetic routes to compounds 3-7.a [hau'H(NCMe)10]4+ [RhI(cod)Cl]2 (i) (vii) [RhH21‘"* "[RhI(cod)(thf)2]1+' (3) [RhIH(n3-TMPP)2]3+ [RhI(nZ-TMPP)(TMPP)(CO)]1+ (4) (7) (iii) (v) (vi) [Rhm(n3-TMPP)(n3-TMPP-0)]2+ [RhI(TMPP)2(CO)2]1+ (5) (6) 2*(i) TMPP (4 equiv), MeCN, 0°C. (ii) NOBF4, MeCN, -40°C or [Cnge][BF4], CH2C12, -40°C. (iii) acetone, r. 15., 24h. (iv) Cp2Co (excess), CO, CH2C12. (v) CO, r. t. (vi) Ar, r. t. (Vii) AgBF4 (2 equiv), THF. (viii) CO, TMPP (2 equiv), 0°C. (ix) [Cp2Fe][BF4] (1 equiv), N2, CH2C12, -15°C. 126 observed the adducts Rh(II)(CO)x (x: 1,2) (although we attempted to do so by in situ infrared or epr studies), their existence is predicated on the basis of the available redox properties of the other intermediates along the reaction pathway. The fact that (33) is too short-lived to be observed by ordinary spectroscopic methods is explained by the very fast electron transfer reaction that takes place between 3 and the strong oxidant (3a) to produce a 50:50 mixture of [Rh(n3—TMPP)2][BF4]3 (4) and [Rh(TMPP)2(CO)2][BF4] (6). Indeed, experiments carried out in NMR tubes under a 12CO atmosphere showed that two diamagnetic products were being produced at the same rate and that they were stable with respect to further reaction - provided the CO atmosphere was maintained. A parallel study using a 50:50 mixture of 12CO/13CO provided solution evidence for the assignment of 6 as a dicarbonyl species. A subsequent X-ray structural study verified this formulation. Apparently electron transfer is faster than initial CO addition to [Rh(n3— TMPP)2]2+, thus, there is never a detectable build-up of a Rh(II)CO+ species and only the diamagnetic products are observed. Under conditions of pumping or purging with an inert gas, [Rh(TMPP)2(CO)2][BF4] (6) readily loses a CO ligand to form [Rh(TMPP)2(CO)2][BF4] (7). However, 7 is unstable with respect to oxidation by the Rh(III) complex, [Rh(n 3- TMPP)2][BF4]3 (4), and a second redox reaction occurs producing [Rh(n3- TMPP)2][BF4]2 (3) and presumably, a Rh(II) monocarbonyl species that readily gives up CO to regenerate 3. The proposed spontaneous redox reaction between 4 and 7 was confirmed by a deliberate 1:1 reaction of the pure compounds as shown in equation 16. Note that no reaction occurs between [Rh(n 3—TMPP)2][BF4]3 and [Rh(TMPP)2(CO)2][BF4 ]. The intermediate Rh(II) carbonyl species 3a is postulated to be a dicarbonyl and not a monocarbonyl complex based on the recognized redox properties of 3 - 7; 127 the monocarbonyl species 7 is known to spontaneously react with [Rh(n3- TMPP)2][BF4]3 (4), therefore the reverse electron transfer process must be unfavorable. I 1+ [RhIII (ns'TMPP)2]% 1 1 (3) (16) TM + (4) [Rh ( PP)2(CO)2L‘ No Reaction (6) The reversible CO chemistry of [Rh(n3-TMPP)2][BF4]2 that has been observed is primarily driven by electronic considerations. That is to say, the relative stabilities of the +1, +2, and +3 oxidation states are controlled by the modulation of electron density at the metal center. The redox reactions of the cycle are driven by the it acceptor strength of CO. The strong n-accepting nature of CO greatly favors the lower oxidation states for which 1:- backbonding is maximized, therefore the effect of CO coordination to 3 is to destabilize the Rh(II) oxidation state relative to Rh(I). Electrochemically, this effect is manifested in a positive shift of the Rh(I)/Rh(II) redox couple, hence the resulting Rh(II) carbonyl complex is more susceptible to reduction. In the case of 3a, the redox couple has moved to well past the Rh(II)/Rh(III) couple for 3. As a result, electron transfer takes place between 3 and 3a to produce [Rh(TMPP)2(CO)2]+ (6) and [Rh(n3-TMPP)2]3+ (4). At that point the reaction is essentially complete, provided a CO atmosphere is maintained. However, upon removal of the CO atmosphere, a coordinated CO ligand is displaced by a TMPP ether group to give [Rh(nZ-TMPP)(TMPP)(CO)]1+. The loss of CO makes the Rh(II) oxidation state more accessible, as evidenced by the less positive oxidation potential for 7 (+ 0.50 V) relative to that of 6 (+ 0.80) (Figure 16). The change in redox properties upon CO dissociation 128 triggers an electron-transfer reaction between [Rh(712-TMPP)(TMPP)(CO)]1+ and [Rh(n3-TMPP)2]3+ that ultimately results in the regeneration of [Rh(n3- TMPP)2][BF4]2 (3). From this, we conclude that while kinetic stabilization is probably an important factor in dictating the course of the reactions with 1:- acceptors ligands, the modulation of the electron density at the metal center (which controls the accessibility of the Rh(I)/Rh(II) and Rh(II)/Rh(III) couple) is also a primary determinant. The preceding analysis provides some insight into ways in which stable paramagnetic adducts of [Rh(n3-TMPP)2][BF4]2 might be prepared. In the current example, the initial disproportionation reaction between 3 and 33 is facilitated by the n-acidity of the incoming ligand. It follows then, that by reducing the n-accepting strength of the incoming ligand, such that the Rh(II)/Rh(I) redox couple occurs at less positive potentials than the Rh(II)/Rh(III) couple of (3), other stable Rh(II) complexes may be isolated. With this in mind, we extended our studies of [Rh(n3-TMPP)2][BF4]2 to include reactions with isocyanides. The weak n-acidity of isocyanide ligands is well documented,16 and as such, they are excellent candidates for preparing stable adducts of [Rh(n3-TMPP)2][BF4]2. The results of these studies are presented in Chapter VI. At this point, we now turn to a discussion of the major influence on the observed chemistry of [Rh(n3-TMPP)2]2+, viz. the flexibility of the ligand. Unlike most other sterically hindered ligands, TMPP allows the metal complex to undergo labilization of weakly held groups and isomerization to more stable geometries. The structures of the trans—Rh(I) complexes 6 and 7 nicely demonstrate this point. It is precisely the lack of structural rigidity that permits complexes of ether-phosphines such as TMPP to undergo reversible small molecule addition reactions. Moreover, the flexibility of the 129 ligand aids in the stabilization of different electronic and structural conformations. As a consequence, complexes of TMPP are expected to exhibit rich redox behavior; a fact that is nicely illustrated by the chemistry of [Rh(n3-TMPP)2][BF4]2. In order to make [RhH(n3-TMPP)2][BF4]2 (3) a viable system for promoting C-C bond formation and other small molecule transformations, disproportionation pathways must be shut down. A strategy for achieving this goal is to tailor the ligand so that it is less accommodating to changes in metal oxidation state and structural conformation but not too rigid to prevent chemistry from occurring at the metal center. A possible solution to this dilemma would be to anchor the TMPP ligand at two points rather than just at the M-P bond, a situation that would prevent free rotation of the phosphine. One such reaction to effect this result is nucleophilic attack on a coordinated methoxy group to produce a stronger metal-phenoxide interaction. An anionic oxygen is not as susceptible to dissociation as an ether substituent, and as a result, the accessibility of the metal center to an incoming substrate is limited. Thus, it is unlikely that substitution chemistry with CO would follow the same pathway as that found in this study. Furthermore, demethylation of a methoxy group to form a phenoxide donor substantially increases the electron density at the metal center. Consequently, this should better stabilize a Rh(II) carbonyl radical with respect to reduction and CO dissociation. Finally, while the overall pathway described in Figure 20 does not represent a simple procedure for reversible CO binding, the reaction between [Rh(TMPP)2(CO)2][BF4] (6) and [Rh(TMPP)2(CO)][BF4] (7 ) is such a process. The high lability of the o-methoxy interaction for [Rh(TMPP)2(CO)][BF4] (7), as evidenced by its solution fluxionality and the facile addition of CO to form 130 6, motivated us to further explore the lability of these weak donor groups. Previously, we observed that solid samples of [Rh(TMPP)2(CO)2][BF4] slowly convert to 7 when exposed to a vacuum for extended periods. This behavior prompted us to examine the reverse process, namely the addition of CO to solid state samples of 7 to form 6. Indeed, powder or Nujol mull samples of [Rh(n3-TMPP)2][BF4]2 (7 ) were observed to reversibly uptake CO at atmospheric pressure by IR spectroscopy. The only impediment to uptake appeared to be the ability of CO to diffuse into the medium. This behavior is- intriguing in the context of CO sensors.1 7 Studies of [Rh(nz- TMPP)(TMPP)CO][BF4] (7) indicate that the complex is unreactive to all other atmospheric gases besides CO. In light of these properties, we set out to investigate the incorporation of [Rh(TMPP)2(CO)][BF4] into porous polymer matrices with the goal of developing a CO sensing material. The results of this endeavor are presented in Chapter V. 10. 131 List of References (a) Trogler, W. C. ed. ”Organometallic Radical Processes"; Elsevier: Amsterdam, 1990. (b) Baird, M. C. Chem. Rev. 1988,88, 1217. (a) Lappert, M. F.; Lendor, P. W. Adv. Organomet. Chem. 1976, 14, 345. (b) Kochi, J. K. "Organometallic Mechanism and Catalysis"; Academic: New York, 1978. (c) Connelly, N. G.; Gieger, W. E. Adv. Organomet. Chem. 1984, 23, 1. (d) Connelly, N. G.; Gieger, W. E. Adv. Organomet. Chem. 1985, 24, 87. (e) Connelly, N. G. Chem. Soc. Rev. 1989,18, 153. (a) Wayland, B. B.; Sherry, A. E.; Coffin, V. L. J. Chem. Soc., Chem. Commun. 1989, 662. (b) Wayland, B. B.; Sherry, A. E. J. Am. Chem. Soc. 1989, 111, 5010. (c) Wayland, B. B.; Sherry, A. E.; Poszmik, G.; Bunn, A. G. J. Am. Chem. Soc. 1992, 114, 1673. (d) Wayland, B. B.; Ba, 8.; Sherry, A. E. J. Am. Chem. Soc. 199], 113, 5305. (e) Sherry, A. E.; Wayland, B. B. J. Am. Chem. Soc. 1990, 112, 1259. (f) Poszmik, G; Wayland, B. B. presented at the 203rd National Meeting of the American Chemical Society, San Francisco, CA, April 5-10, 1992; paper INOR 67. (a) Bennett, M. A.; Lonstaff, P. A. J. Am. Chem. Soc., 1969,91, 6266. (b) Vleck, A. Inorg. Chim. Acta 1980, 43, 35. (c) Holah, D. G.; Hughes, A. N.; Hui, B. C. Can. J. Chem. 1975,53, 3669. (d) Valentini, G.; Braca, G.; Sbrana, G.; Colligiani, A. Inorg. Chim. Acta 1983, 69, 215. (e) Valentini, G.; Braca, G.; Sbrana, G.; Colligiani, A. Inorg. Chim. Acta 1983, 69, 221. Empsall, H. D.; Hyde, E. M.; Jones, C. E.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1974, 1980. (a) Mason, R.; Thomas, K. M.; Empsall, H. D.; Fletcher, S. R.; Heys, P. N.; Hyde, E. M. Jones, C. E; Shaw, B. L. J. Chem. Soc., Chem. Comm. 1974, 612. (a) Empsall, H. D.; Hyde, E. M.; Shaw, R. L. J. Chem. Soc., Dalton Trans. 1975, 1690. (b) Empsall, H. D.; Heys, P. N.; McDonald, W. S.; Norton, M. C.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1978, 1119. Gray, H. B.; Hendrickson, D. N.; Sohn, Y. S. Inorg. Chem. 197], 10, 1559. SDP: Structure Determination Package, Enraf-Nonius, Delft, The Netherlands, 1979. TEXSAN - TEXRAY Structure Analysis Package, Molecular Structure Corporation, 1985. 11. 12. 13. 14. 15. 16. 17. 132 (a) MITHRIL: Integrated Direct Methods Computer Program, Gilmore, C. J. J. Appl. Cryst. 1984, 17, 42. (b) DIRDIF: Direct Methods for Difference Structure, An Automatic Procedure for Phase Extension; Refinement of Difference Structure Factors. Beurskens, R. T. Technical Report 1984. For evidence supporting similar exchange processes see (a) El-Amouri, H.; Baltsoun, A. A.; Osborn, J. A. Polyhedron 1988, 7, 2035. (b) English, A. D.; Meakin, P.; Jesson, J. P. J. Am. Chem. Soc. 1976, 98, 7590 (c) Volger, H. C.; Vrieze, K.; Praat, A. P. J. Organomet. Chem. 1968, 14, 429. Bader, A.; Lindner, E. Coord. Chem. Rev. 199], 108, 27. and references therein. Treichel, P. M.; Mueh, H. J.; Bursten, B. E. Isr. J. Chem. 1976/77, 15, 253. For instance see (a) Dunbar, K. R.; Haefner, S. C.; Burzynski, D. J. Organometallics 1990, 9, 1347. (b) Jeffrey, J. C.; Rauchfuss, T. B. Inorg. Chem. 1979, 18, 2658. (c) Granziani, R.; Bombiere, G.; Volponi, L.; Panattoni, C.; Clark, R. J. H. J. Chem. Soc. (A) 1969, 1236. ((1) De V. Steyn, M. M.; English, R. B.; Ashworth, T. V.; Singleton, E. J. Chem. Res. (S) 198], 267. Treichel, P. M. Adv Organomet. Chem. 1973, 11 , 21. Biosensors and Chemical Sensors: Optimizing Performance Through Polymeric Materials, ACS Symposium Series 487; Edelman, P. G.; Wang, J. Eds., American Chemical Society: Washington, D. C., 1992. CHAPTER V REVERSIBLE CARBON MONOXIDE ADDITION TO SOL-GEL DERIVED COMPOSITE FILMS CONTAINING THE MOLECULE [Rh(TMPP)2(CO)]“ 134 1. Introduction The reversible solution chemistry of the novel mononuclear Rh(II) complex [Rh(n3-TMPP)2][BF4]2 (TMPP = tris(2,4,6-trimethoxyphenyl)- phosphine) with carbon monoxide was presented in Chapter 4. The reaction was found to proceed by a redox pathway that involves the formation of the Rh(I) carbonyl intermediates [Rh(TMPP)2(CO)][BF4] (7 ) and [Rh(TMPP)2(CO)2][BF4] (6); these complexes were fully characterized by X- ray crystallography, cyclic voltammetry, as well as by infrared and NMR spectroscopies. In solution, complex 7 rapidly and reversibly binds carbon monoxide under ambient conditions to form the dicarbonyl species 6 (eq 1). In the absence of a CO atmosphere, 6 loses CO to reform 7. The facile nature of the carbon monoxide addition to [Rh(TMPP)2(CO)][BF4] (7) is exemplified by the observation that finely divided powder and Nujol mull samples of 7 are also capable of reversibly uptaking CO. Not surprisingly, these reaction Me Me- Me 1 o’ “I 1* Me- O ‘ArArTI 131-Ar 15‘ + C: 0 Me— 0 P‘ OC—lith— co (17) Me’ O‘Iid’l— CO ‘7— Arv\\\ E‘AI‘I Ar' \" E‘ ' - CO Ar' Ar' OMe [Rh(TMPP)2(CO)]1+ (7) Ar‘ = GOMe [Rh(TMPP)2(CO)2]1+ <6) OMe rates are sluggish due to poor diffusion of CO into the solid; this situation prompted us to investigate the incorporation of 7 into a porous material that would trap the molecular cationic species yet facilitate diffusion of CO into the matrix. Low temperature sol-gel techniques can be used to create porous 135 oxide glasses (pore size < 100 A) that encapsulate large complex molecules yet allow diffusion of small substrates into the matrix.1 Such oxide galsses fromed by the hydrolysis and polycondensation of metal alkoxides are optically transparent and can be readily probed by a number of spectroscopic techniques. Previous work has demonstrated the general usefulness of SOI- gel derived glasses for the immobilization of guest molecules such as inorganic clusters, porphyrins, and lanthanide cryptate complexes.” Recently, the process has been adapted for the incorporation of biochemically active molecules that may lead to the development of novel biosensing materials.4 The use of sol-gel techniques provides for the immobilization of a sensing molecule, in this case [Rh(TMPP)2(CO)][BF4] (7), in an environment that can be probed both by spectroscopic and electrochemical techniques, while the porous nature of the films permits facile diffusion of an exogenous substrate, such as CO, that can then bind to the sensing molecule. Herein we report the synthesis and CO binding properties of zirconia and titania glasses impregnated with a rhodium ether-phosphine compound, the results of which are promising for future adaptations of such molecular species as chemical sensors. The work described in this chapter was performed in collaboration with Professor Kris A. Berglund and Dr. Joel I. Dulebohn in the Departments of Chemical and Agricultural Engineering, Michigan State University. 136 2. Experimental A. Preparation of Composite Films Containing [Rh(TMPP)2(C0)] [BF4] (7) (l) Zirconia Films An amount of lauric acid (0.161 g, 0.80 mmol) was added to 1 mL of 70% Zr(OPrn)4 in propanol (2.23 mmol) in a glass vial. The mixture was placed in a sonicator (Bransonic 220) until the lauric acid had completely dissolved. To this solution, was added an amount of valeric acid (1.7 mL, 15.6 mmol) and 0.2 mL of water. A solution of [Rh(TMPP)2(CO)][BF4] (7 ) (5.9 mg, 4.6 x 10‘3 mmol) in 50 uL of CH2C12 was added to 1 mL of the zirconia film solution. The resulting yellow solution was filtered through a 0.22 mm Millex-GS millipore filter. (2) Titania Films A quantity of of lauric acid (0.25 g, 1.25 mmol) was dissolved in neat Ti(OPri)4 (1 mL, 3.36 mmol). The solution was sonicated to ensure that the lauric acid was completely dissolved. A mixture of valeric acid (2.5 mL, 23 mmol) and water (0.3 mL) were then added to the solution. A solution of [Rh(TMPP)2(CO)][BF4] (7) (0.0059 g, 4.6 x 10'3 mmol) in 50 uL of CH2C12 was pipetted into 1 mL of the titania film solution. The resulting yellow solution was filtered through a 0.22-mm Millex-GS millipore filter to remove unreacted 7. B. Flow Rate Determination A schematic diagram for the apparatus used to estimate CO concentrations in the films is shown in Figure 22. Flow rates of CO and the argon carrier gas were controlled by adjustment of screw clamps connected to each line and were estimated by measuring the time required for a soap bubble to travel a predetermined distance through a glass pipette. The CO 137 glass _ pipette soap reservoir Ar _, Fl to IR cell —> l\| [J switch —> screw clamp O O l :l<— [22 Figure 22. Schematic diagram of the apparatus used to estimate CO concentrations. 138 flow rate was adjusted and measured, first in the absence of argon, then the argon flow rate was adjusted and the combined flow rate for the two gases was measured. The flow rates for CO and the combined gases were measured five times and the average was determined. The % CO concentration was taken as the inverse of the flow rate ratios. C. Electrochemical Measurements Films used for electrochemical studies were prepared as described in section A. However, in order to make the films electrochemically active, 2.0 mg of [Li][CF3SO3] (0.013 mmol) was added to 0.1 mL of the film solution as a supporting electrolyte. The films were cast onto a Pt disk electrode using a cotton swab and dried in a stream of cool air. Cyclic voltammetry measurements were performed in 0.1 M [Li][CF3803] / H20 at a scan rate of 200 mV/s using a platinum disk working electrode and a Ag/AgCl reference electrode. 3. Results Sol-gel derived titania and zirconia composite films containing [Rh(TMPP)2(CO)l[BF4] (7 ) were prepared by addition of a dichloromethane solution of 7 to titania and zirconia film solutions. Thin films were fabricated by either spin casting for 5 minutes or by application to the substrate with a cotton swab followed by drying in a stream of cool air. Infrared spectroscopic studies on the films were performed by casting the composite zirconia or titania films onto a germanium ATR crystal and placing the crystal in an liquid ATR cell holder. Initially, a single carbonyl stretch was seen at 1975 cm‘l, which corresponds to the cation species [Rh(TMPP)2(CO)]+ (7) (v(CO)(3}{2(312 = 1970 cm'l) immobilized in the film. Upon exposure of the film to a CO atmosphere, the band at 1975 cm‘1 disappeared and a second, more 139 94 — J o \4 O l: 53 _ E a 92 - (a) g 1975 cm-1 5° —1 ‘ (b) ‘ 2013 cm'1 90 _ I l 1 l I 17 2000 1800 wavelength cm-1 Figure 23. Infrared spectra of a zirconia composite film of [Rh(TMPP)2(CO)][BF4] (7): (a) in the absence of CO. (b) exposed to CO. 140 intense, band appeared at 2013 cm'l, signaling the binding of CO to 7 to yield the dicarbonyl species [Rh(TMPP)2(CO)2][BF4] (6) (V(CO)CH2012 = 2011 cm-1) (Figure 23). The absence of the 1975 cm'1 band indicates that the conversion of 7 to 6 within the film is quantitative. With regards to sensitivity, the presence of CO was observable in gas mixtures of CO and Ar containing less than 0.2% CO, which was the lowest measurable concentration based on our present method for determining flow rates. In the lower concentration range, the conversion of 7 to 6 is less than quantitative as evidenced by the continued presence of the CO band for 7 at 1975 cm'l. Removal of CO from the system was effected by purging the IR cell with argon for one minute, whereupon the band at 2013 cm'1 diminished with concomitant reappearance of the 1975 cm'1 band; this observation is consistent with complete conversion of [Rh(TMPP)2(CO)2][BF4] (6) back to the parent molecular species [Rh(TMPP)2(CO)][BF4l (7). Alternatively, the loss of CO from the material may be achieved by simply exposing the film to the atmosphere, although the process is significantly slower under these conditions. As judged by the invariance of the IR spectra, zirconia and titania composite films of [Rh(TMPP)2(CO)][BF4] (7 ) are selective for CO in the presence of other gases such as 02, C02, N2, and H2. In addition to infrared spectroscopy, the reversible addition of CO to [Rh(TMPP)2(CO)][BF 4] (7 ) within the composite films was monitored by electronic spectroscopy. A zirconia composite film containing 7 was spin cast onto a quartz slide and placed in a 1 cm quartz cuvette capped with a rubber septum. The cell was purged for 15 min with CO giving rise to an electronic spectral feature at 435 nm which is comparable to that seen in solution for [Rh(TMPP)2(CO)2][BF4] (6) (AmapozClz = 438 nm) (Figure 24). Subsequent removal of CO from the cell by purging with Ar resulted in the disappearance Abs 141 (a) [Rh(FMPP)2(C0)llBF4l 438 run I (b) [Rh(TMPPh(C0)2][BF4l 300 Abs Figure 24. foo 560 E00 wavelength (nm) 435nm H (b) r :r‘ :1. w; (a) I I I I I F r I I r I 400 500 wavelength (nm) Electronic absorption spectra of (a) [Rh(TMPP)2(CO)][BF4] and (b) [Rh(TMPP)2(CO)2][BF4] in: (I) CH2C12. (II) a zirconia composite film. 142 of the transition at 435 nm, leading to the conclusion that 6 had been reconverted to the monocarbonyl derivative 7. Initial electrochemical studies of the sol-gel derived composite films containing [Rh(TMPP)2(CO)][BF4] (7 ) were carried out by casting a zirconia film onto a platinum disk electrode. Unfortunately, these films were found to be electrochemically inactive as no current response was observed. To circumvent this problem [Li][CF3SO3] was added to the film solution as a supporting electrolyte prior to casting onto the electrode surface. A cyclic voltammogram of the composite film in 0.1 M [Li][CF3SO3]/ H2O exhibited an essentially featureless voltammogram (Figure 25). Upon purging the electrochemical cell with CO for a brief period of time (< 30 sec), an irreversible oxidation wave appears at Ema: + 0.74 V with a chemical return wave located at EM: + 0.31 V vs. Ag/AgCl. After purging with N2 for several minutes, the oxidation wave gradually disappears and the original cyclic voltammogram is obtained. Purging the solution with N2 removes CO from the system and results in re-formation of 7, in agreement with the reversible addition of CO that was observed spectroscopically. Films containing only [Li][CF3S03], when exposed to CO under identical conditions, exhibited no significant change in current response, indicating to us that the observed signal is due to the encapsulated complex bound to CO. Although the electrochemical behavior of the carbonyl complexes 7 and 6 incorporated in these composite films differs from the solution behavior, it is apparent that CO is binding to the rhodium center; more importantly, the process is reversible. Such behavior is in contrast to the irreversible CO chemistry reported for conducting polymer films containing ferrocenylferraazetine.5 In order to make electrochemical detection a more viable means of sensing CO in the present case, the use of a more conducting matrix is in order. One 143 (a) Volts vs. Ag/AgCl Figure 25. Cyclic voltammograms of a zirconia composite film containing [Rh(TMPP)2(CO)][BF4] (7) and LiCFgSO3 cast onto a platinum disk electrode: (a) under a N2 atmosphere. (b) after purging the cell for 30 seconds with CO. (c)-(e) after flushing the CO saturated cell with N2 for 2, 5 and 10 min. 144 potential solution is the incorporation of [Rh(TMPP)2(CO)][BF4] (7 ) into solid ionic conducting polymer films such as MEEP (poly[bis(2-(2- methoxy)ethoxy)phosphazene]).6 4. Discussion These studies demonstrate that the molecular cationic complex [Rh(TMPP)2(CO)]1+ (7) reversibly binds CO within a glassy polymer matrix to form the dicarbonyl species [Rh(TMPP)2(CO)2]1+ (6). The facile nature of the chemistry is a direct consequence of an exceedingly labile metal-ether bond, which facilitates substitution reactions even in the solid state. Furthermore, the metal-ether interaction appears to be selective towards CO as other atmospheric gases do not react. Another advantage of this system is that the substrate binds directly to the metal center. As a result, the addition of CO conveniently gives rise to dramatic changes in the spectroscopic and redox properties of the rhodium cation, commensurate with coordination of a strong niacceptor to the metal center. Because of the transparent characteristcs of the sol-gel derived films, these spectral and electrochemical changes are readily detected. In spite of these attributes, a number of concerns must be addressed before the application of these materials as sensing devices can be realized; these include long term stability of material, conditions under which the films will operate, detection limits and response times. Nevertheless, on the basis of these preliminary results with titania and zirconia films incorporating [Rh(TMPP)2(CO)][BF 4] (7 ), the development of other CO sensing films using molecular composites of ether-phosphine complexes appears attractive. 145 List of References (a) Brinker, C. J .; Scherer, G. Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing; Academic Press: San Diego, 1989. (b) Sanchez, C.; Livage, J. New J. Chem. 1990,14, 513. (a) Dulebohn, J. 1.; Van Vlierberge, B.; Berglund. K. A.; Lessard, R. B.; Yu, J .; Nocera, D. G. Mat. Res. Soc. Symp. Proc. 1990, 180, 733. (b) Lessard, R. B.; Wallace, M. M.; Oertling, W. A.; Chang, C. K.; Berglund. K. A.; Nocera, D. G. Mat. Res. Soc. Symp. Proc. 1989, 155, 109. (c) Lessard, R. B.; Berglund. K. A.; Nocera, D. G. Mat. Res. Soc. Symp. Proc. 1989, 155, 119. (d) Newsham, M. D.; Cerreta, M. K.; Berglund. K. A.; Nocera, D. G. Mat. Res. Soc. Symp. Proc. 1988, 121, 627. For other recent examples see: (a) Kuselman, I.; Kuyavskaya, B. I.; Lev, O. Analytica Chimica Acta 1992, 256, 65. (b) Dunn, B.; Zink, J. I. J. Mater. Chem. 199], 1, 903. (c) McKiernan, J. M.; Yamanaka, S. A.; Knobbe, E.; Pouxviel, J. C.; Parveneh, S.; Dunn, B.; Zink, J. I. J. Inorg. Organomet. Polym. 1991,], 87. (d) Schwok, A.; Avnir, D.; Ottolenghi, M. J. Am. Chem. Soc. 199], 113, 3984 and references therein. (e) Haruvy, Y.; Webber, S. E. Chem. Mater. 1991, 3, 501 and references therein. (f) Zusman, R.; Rottman, C.; Ottolenghi, M.; Avnir, D. J. Non- cryst. Solids 1990, 122, 107. (h) McKiernan, J. M.; Yamanaka, S. A.; Dunn, B.; Zink, J. I. J. Phys. Chem. 1990, 94, 5652. (i) McKiernan, J. M.; Pouxviel, J. C.; Dunn, B.; Zink, J. I. J. Phys. Chem. 1989, 93, 2129. (j) Pouxviel, J. C.; Dunn, B.; Zink, J. I. J. Phys. Chem. 1989, 93, 2134. (k) Durm, B.; Knobbe, E.; McKiernan, J. M.; Pouxviel, J. C.; Zink, J. 1. Mat. Res. Soc. Symp. Proc. 1988, 121, 331. (a) Ellerby, L. M.; Nishida, C. R.; Nishida, F.; Yamanaka, S. A.; Dunn, B.; Selverstone-Valentine, J .; Zink, J. 1. Science 1992, 255, 1113. (b) Yamanaka, S. A.; Nishida, F.; Ellerby, L. M.; Nishida, C. R.; Dunn, B.; Selverstone-Valentine, J .; Zink, J. 1. Chem. Mater. 1992,4, 495. (c) Slama-Schwok, A.; Ottolenghi, M.; Avnir, D. Nature 1992,355, 240. (a) Mirkin, C. A.; Wrighton, M. S. J. Am. Chem. Soc. 1990, 112, 8596. (b) Mirkin, C. A.; Valentine, J. R.; Ofer, D.; Hickman, J. J .; Wrighton, M. S. In Biosensors and Chemical Sensors: Optimizing Performance Through Polymeric Materials, ACS Symposium Series 487 ; Edelman, P. G.; Wang, J. Eds., American Chemical Society: Washington, D. C., 1992, chapter 17, p. 218. (a) Allcock, H. R.; Austin, P. E.; Neenan, T. X.; Sisko, J. T.; Blonsky, P. M.; Shriver, D. F. Macromolecules 1986, 19, 1508. (b) Blonsky, P. M.; Shriver, D. F.; Austin, P. E.; Allcock, H. R. Solid State Ionics 1986, 18, 258. (c) Blonsky, P. M.; Shriver, D. F.; Austin, P. E.; Allcock, H. R. J. Am. Chem. Soc. 1984,106, 6854. (d) Austin, P. E.; Riding, G. J.; Allcock, H. R. Macromolecules 1983,16, 719. CHAPTER VI CHEMISTRY OF [Rh11(n3-TMPP)2] [BF4]2 WITH ISOCYANIDE LIGANDS 146 147 1. Introduction A survey of the literature reveals that most of the documented research involving mononuclear rhodium complexes has focused on the +1 and +3 oxidation states, due, in large part, to their demonstrated involvement in homogeneous catalytic processes.1 Only a limited number of reports have addressed the coordination chemistry of paramagnetic Rh(II),2 and even fewer have focused on mononuclear organometallic complexes of Rh(II).3 The unusual stability of 3 has presented us with a rare opportunity to investigate these elusive classes of compounds.4 In light of the recent reports by Wayland and co—workers of carbon monoxide and methane activation by Rh(II) metalloradicals,5 this is a particularly attractive area of research. In Chapter IV, we demonstrated that [Rh(n3-TMPP)2][BF4]2 (3) reacts reversibly with CO through a pathway that involves the formation of Rh(I) carbonyl and Rh(III) intermediates. From the chemistry of 3 with CO it was apparent to us that the primary factor promoting the redox chemistry was the it-accepting ability of the incoming ligand. Intrigued'by this unusual chemistry, we set out to explore the reactivity of [Rh(n3-TMPP)2]2+ (3) with other n-acceptors with the goal of understanding the factors that influence the stability of Rh(II) complexes. Herein, we describe the synthesis and full characterization of a series of novel square planar Rh(II) complexes stabilized by isocyanides and phosphines. 2. Experimental A. Synthesis All reactions were carried out under an argon atmosphere by the use of standard Schlenk-line techniques unless otherwise stated. The starting materials tris(2,4,6-trimethoxyphenyl)phosphine (1 ) and [Rh(n3- TMPP)2][BF4]2 (3) were prepared as described in chapters II and III 148 respectively. The reagents tert-butyl isocyanide, iso-propyl isocyanide, n- butyl isocyanide, cyclohexyl isocyanide and cobaltocene were purchased from Strem Chemicals and used without further purification. Methyl isocyanide was prepared according to a literature procedure.6 (1) Preparation of [Rh(TMPP)2(CNBut)2][BF4]2 (9) An amount of ButNC (34 uL, 0.300 mmol) was carefully syringed into a rapidly stirring solution of [Rh(n3-TMPP)2][BF4]2 (3) (0.200 g, 0.150 mmol) in CH2C12 (3 mL). The solution color immediately changed from purple to black/purple. After stirring for 10 min, the solution was layered with 20 mL of THF. After diffusion had occurred, dark purple crystals separated from the solution; these were collected by suction filtration in air, washed with THF (3 x 5 mL) and Et20 (2 x 5 mL), and subsequently dried under vacuum; yield, 0.213 g (90%). In an alternative work-up procedure, the solution was filtered in air after 10 min of reaction and diluted with CH2C12 (20 mL) followed by THF (8 mL). Reduction of the solution volume to 3 - 5 mL on a rotary evaporator produces a purple crystalline product upon standing overmght. Anal. Calcd for C64H74F8P2013N2B2Rh: C, 50.98; H, 4.95. Found: C, 50.45; H, 5.88. IR (CH2C12, cm'l): v(CEN), 2200 vs; (Nujol mull, CsI, cm-l): v(CEN), 2198 vs; other, 1595 vs, 1578 vs, 1411 s, 1332 s, 1294 w, 1228 s, 1207 s, 1185 w, 1160 s, 1122 s, 1087 s, 1054 s, 1026 s, 950 m, 917 m, 815 m, 675 w, 640 w, 536 w, 522 w, 479 m, 442 w. Electronic absorption spectrum (CHZCIZ) kmax, nm (e): 819 nm (1770), 546 (630), 319 (22,400), 257 (56,500). 1H NMR and 31P NMR (CD2C12) 8 ppm: not observed. Cyclic voltammogram (0.1 M TBABF4/CH2C12, vs Ag/AgCl): E1/2(red) = - 0.04 V. (2) Preparation of [Rh(TMPP)2(CNPri)2][BF4]2 (10) Dropwise addition of a solution of PriNC (13.6 uL, 0.150 mmol) in 7 mL of CH2C12 to a solution of [Rh(n3-TMPP)2][BF4]2 (3) (0.100 g,, 0.075 mmol) in O 149 3 mL of CH2C12 produced a black/purple solution over a period of 10 min. The reaction was stirred for an additional 10 min, after which time 5 mL of CH2012 and 10 mL of THF were added. The solution was concentrated to a volume between 5 - 8 mL on a rotary evaporator and filtered through Celite. Upon standing overnight, a crop of purple crystals formed; these were isolated and washed with copious amounts of THF and EtZO; yield 0.053 g (48%). Anal. Calcd for C62H80F8P2013N2B2Rh: C, 50.32; H, 5.45. Found: C, 49.54; H, 5.50. IR (CH2C12, cm‘l): v(CEN), 2211 vs; (Nujol mull, CsI, cm‘l): v(CEN), 2209 vs; other, 1595 vs, 1577 vs, 1411 s, 1331 s, 1297 w, 1231 s, 1206 s, 1184 w, 1161 s, 1125 s, 1088 s, 1054 s, 1025 s, 949 m, 917 m, 812 m, 478 m. Electronic absorption spectrum (CH2C12) Amax, nm (a): 843 nm (1847), 558 (645), 318 (20,685), 258 (57,933). Cyclic voltammogram (0.1 M TBABF4/CH2C12, vs Ag/AgCl): E1]2(red) = + 0.01 V. (3) Reaction of [Rh(n3-TMPP)2] [BF4]2 (3) with other isocyanides (i) Methyl isocyanide: (a) one equivalent. In a typical reaction, MeNC (2 uL) was added to a solution of [Rh(n3-TMPP)2][BF4]2 (3) (0.050 g, 0.037 mmol) in 5 mL of CH2CIZ. The solution immediately turned from purple to red with the concominant precipitation of a small quantity of red solid (see below). An IR monitoring study of the solution reveals the immediate formation of several CEN stretching vibrations; v(CEN) cm'1: 2239 s, 2248 s, 2267 m, 2176. After 10 min, an additional band appeared at 2150 cm'1. (b) five equivalents. To a solution of [Rh(n 3-TMPP)2][BF4]2 (3) (0.100 g, 0.75 mmol) in 5 mL of CH2C12 was added 5 equivalents of MeNC (20 uL, 3.75 mmol). A red flocculent solid precipitated from the red/orange solution. The resulting suspension was stirred for 5 min, after which time the solid was collected by filtration, washed with a minimal amount of CH2C12 (< 5 mL) and THF ( 2 x 5 mL) and dried under vacuum; yield: 0.43 g. IR (Nujol, CsI) 150 cm'1: v(CEN) 2238 s; [BF4]' 1050 br. 1H NMR (CD3CN) 8 ppm: 3.41 (s, 12H), 3.57(s, 18H, o-OCH3), 3.83 (s, 9H, p-OCH3), 6.21 (t, lJP-H = 2.1 Hz, 6H, m-H). 31P NMR (CD3CN) 5 ppm: - 39.5 (t, J = 42.7 Hz). Cyclic voltammogram (0.1 M TBABF4/CH2C12, vs Ag/AgCl): Ep,c = - 0.46 V, Ep,c = - 1.71 V. (ii) Cyclohexyl isocyanide. To a solution of [Rh(n 3-TMPP)2][BF4]2 (3) (0.050 g, 0.037 mmol) in 5 mL of CH2C12 was added 2 equivalents of CNCy (9.3 uL, 0.075 mmol). The purple solution immediately turned dark purple and then finally black. The IR spectrum measured within the first minute exhibited three bands; v(CEN): 2245 vw, 2210 m, 2173 m. Over a period of several hours, the low energy band continued to grow in as the solution color became pale. (iii) n-Butyl isocyanide. In a typical reaction, two equivalents of CNBun (7.8 uL, 0.075 mmol) were added to a purple solution of [Rh(n3- TMPP)2][BF4]2 (3) in 5 mL of CH2Cl2. The solution immediately became dark brown. After 30 min, the solution IR spectrum showed several CEN stretching vibrations; v(CEN); 2292 w, 2255 vs, 2220 w, 2160. The solvent was removed under reduced pressure and the resulting residue was redissolved in THF (5 mL), filtered in air to remove unreacted 3 and finally pumped to a residue. (4) Reactions of [Rh(TMPP)2(CNBut)2] [BF4]2 (9) (i) Cobaltocene: Preparation of [Rh(TMPP)2(CNBut)2][BF4] (11). A solution containing [Rh(TMPP)2(CNBut)2][BF4]2 (9) (0.100 g, 0.066 mmol) and 1 equivalent of Cp2Co ( 0.013 g, 0.066 mmol) in 5 mL of CH2C12 was stirred at r. t. for 30 min. The yellow solution was evaporated to dryness, and the resulting residue was redissolved in THF (10 mL). The solution was filtered through a Celite plug to remove undissolved [Cp2Co][BF4] and reduced in volume to 5 mL. Diethyl ether (15 mL) was added and the m 151 solution was again filtered. The solvent was then removed under vacuum and the yellow residue was taken up in 5 mL of CH2C12. To this solution, 35 mL of diethyl ether was added creating a supersaturated solution from which a yellow crystalline solid separated. The crystals were collected by filtration and washed with diethyl ether (2 x 5 mL); yield, 0.060 g (64 %). IR (CH2C12, cm‘l): v(CEN), 2118 vs; (Nujol mull, CsI, cm'l): v(CEN), 2198 vs; other, 1595 vs, 1578 vs, 1411 s, 1332 s, 1294 w, 1228 s, 1207 s, 1185 w, 1160 s, 1122 s, 1087 s, 1054 s, 1026 s, 950 m, 917 m, 815 m, 675 w, 640 w, 536 w, 522 w, 479 4 m, 442 w. Electronic absorption spectrum (CH2C12) km”, nm: 395. 1H NMR (CD2C12, 25°C) 5 ppm: -C(CH3)3 0.86 (s, 9H), 0.88 (s, 9H); ~0CH3 3.07 (s, 6H), 3.16 (s, 6H), 3.30 (s, 6H), 3.37 (s, 6H), 3.53 (s, 6H), 3.70 (s, 6H), 3.73 (s, 6H), 3.77 (s, 6H), 3.83 (s, 3H), 3.85 (s, 3H); meta-H 5.88 (br, 8H), 6.02 (br, 2H), 6.12 (br, 2H). 31P NMR (CDZClz, 25°C) 8 ppm: PA, -19.1 (d, 1JRh_p = 129.7 Hz); PB , -20.2 (d, 1JRh_p = 129.7 Hz). Cyclic voltammogram (0.1 M TBABF4/CH2C12, vs Ag/AgCl): El/2(ox) = - 0.04 V. (ii) TMPP. A mixture of [Rh(TMPP)2(CNBut)2][BF4]2 (9) (0.020 g, 0.013 mmol) and TMPP (0.007 g, 0.013 mmol) was dissolved in approx. 0.7 mL of CD3CN. The resulting purple solution was transferred to a NMR tube. No reaction had occurred after 12 h at r.t. as evidenced by 1H NMR spectroscopy. (iii) tert-Butyl isocyanide. A solution of [Rh(TMPP)2(CNBut)2][BF4]2 (9) (0.050 g, 0.033 mmol) in 5 mL of CH2C12 was treated with 3.8 uL of CNBut (0.033 mmol) and stirred at r.t. After 10 min, the solution color had gradually changed from purple to black/green and eventually to green. After 12 h an infrared spectrum of the solution was obtained; IR (CH2C12) cm'l: 2167 s, 2118 m, 2230 w. An analogous 31P NMR study in CD3CN revealed the presence of unligated TMPP and [TMPP-CH3]+ in addition to a new species; 31P NMR (CD3CN) 8 ppm: - 27.5 (d, lJRh-P = 77.8 Hz) 152 (iv) Carbon monoxide. An amount of [Rh(TMPP)2(CNBut)2][BF4]2 (9) (0.025, 0.017 mmol) was dissolved in 5 mL of CH2C12. The solution was purged with CO for 30 min. An infrared spectrum of the solution showed that no reaction had occurred under these conditions. B. X-ray Crystallography The structures of [Rh(TMPP)2(CNBut)2][BPh4]2 and [Rh(TMPP)2(CNBut)2][BF4] were determined by application of general procedures which have been fully described elsewhere.7 Geometric and intensity data were collected on a Rigaku AFC6S diffractometer with graphite monochromated MoKoc (it-a = 0.71069 A) radiation and were corrected for Lorentz and polarization effects. Important crystallographic data are summarized in Table 13. All calculations were performed with the use of VAX computers on a cluster network within the Department of Chemistry at Michigan State University by using the Texsan software package of the Molecular Structure Corporation.8 (1) [Rh(TMPP)2(CNBut)2][BPh4]2 (9) (i) Data Collection and Reduction. We were unable to obtain suitable single crystals of [Rh(TMPP)2(CNBu")2]2+ as a [BF4]‘ salt for X-ray diffraction studies. However, metathesis of [BF4]' with KIBPh4] in acetone followed by slow diffusion of Eth yielded large dark purple crystals of [Rh(TMPP)2(CNBut)2][BPh4]2. A block-shaped crystal of approximate dimensions 0.26 x 0.26 x 0.29 mm3 was selected and secured to the tip of a glass fiber with epoxy cement. Cell constants were obtained from a least- squares fit of 25 centered reflections in the range 15 < 20 < 20° and were consistent with a monoclinic cell. Intensity data were gathered at 23 i 1°C in the range 4 S 29 S 47° using the (1)-scan technique. Reflections with I < 106(1) Table 13. Summary of crystallographic data for [RhH(TMPP)2(CNBut)2] [BPh4]2 (9a). Formula RhP20180112H124B2 Formula weight 1972.69 Space group P21/c a, A 26.810(7) b, A 14.076(2) c, A 27.809(6) a, deg 90 [3, deg 101.35(2) 7, deg 90 V, A3 10,289(4) Z 4 dcalc, g/cm3 1.273 )1 (Mo Kat), cm'1 2.57 Temperature, °C 23:2 °C Ra 0.056 wa 0.071 Quality-of-fit indicatorc 3.80 312:: | lFol- chll/ztrol bRw = [2w( I F0 I— IFc l)2/>:w To I211/2; w = 1/02( lFo l) CQuality-of-fit = [2w( IFo | — ch DZ/(Nobs-Nparametersfl 10- 154 were rescanned a maximum of three times and the counts were accumulated to ensure good counting statistics. Three representative reflections were monitored at regular intervals and exhibited a 1.5% loss in intensity. A linear correction factor was applied to the data to account for the observed decay in intensity. After equivalent data were averaged, there remained 9109 data with F02 2 36(Fo)2 which were used in the structure solution and refinement. (ii) Structure Solution and Refinement. The position of the rhodium~ atom was found by direct methods.9 The remaining non-hydrogen atoms were located and refined through successive least-squares cycles and difference Fourier maps. After isotropic convergence, an empirical absorption correction was applied using the program DIFABS.10 Thermal parameters for all non-hydrogen atoms were then refined anisotropically with the exception of the three methyl carbons C(62), C(63) and C(64) on one of the isocyanide ligands; these were treated isotropically. Hydrogen atoms were calculated at fixed positions and were not refined. Anisotropic refinement of 1219 parameters converged to give R = 0.056 and R“. = 0.071. The quality-of- fit index was 3.80 and the largest shift/esd = 0.32. 3. Results A. Synthesis and Spectroscopic Characterization of [Rh(TMPP)2(CNR)2] [BF4]2 (R = But, Pri). Dichloromethane solutions of [Rh(n3-TMPP)2][BF4]2 (3) react smoothly with two equivalents of CNR (R = But, Pri) to yield dark purple solutions that exhibit a characteristic v(C_=.N ) band at 2200 cm'1 for CNBut and 2211 cm'1 for CNPri. Both values appear at a higher energy than that of the free isocyanide which is indicative of a higher valent metal complex in which the _r_‘-7" ' ~. .4 ._. PM“: ,. :‘rrv'm'l 155 isocyanides act primarily as a-donors.11 Reaction with only one equivalent of isocyanide results in a 1:1 mixture of product and starting material, consistent with the formulation of the stable product as a bis adduct of (3). In both cases, concentration of the reaction solution followed by addition of THF produces a dark purple microcrystalline solid. The products, [Rh(TMPP)2(CNBut)2][BF4]2 (9) and [Rh(TMPP)2(CNPri)2][BF4]2 (10), were isolated in 90% and 48% yield, respectively. In spite of the highly reactive nature of many radical d7 rhodium species, these Rh(II) di-isocyanide complexes are remarkably air-stable both in the solid state and in solution. The electronic spectra of [Rh(TMPP)2(CNBut)2][BF4]2 (9) and [Rh(TMPP)2(CNPri)2][BF4]2 (10) both exhibit two prominent low energy bands in the visible region at 819 nm (e = 1770 M‘lcm'l) and 546 nm (e = 630 M‘lcm'l) for 9 and at 843 nm (e = 1847 M' 1cm‘l) and 558 nm (e = 645 M' 1cm'l) for 10; these transitions give rise to the characteristic dark purple color exhibited by these paramagnetic complexes. Although 9 and 10 are stable with respect to oxygen and moisture, both complexes readily react with additional isocyanide to give a mixture of diamagnetic products. NMR spectral measurements of CD3CN solutions of [Rh(TMPP)2(CNBut)2][BF4]2 (9 ) in the presence of excess tert-butyl isocyanide showed that TMPP dissociates during the reaction to form a new Rh(TMPP) species. Based on the symmetrical appearance of the proton resonances, this compound is likely a Rh(I) monophosphine complex with the general formula [Rh(TMPP)(CNBut)x]1+ (x = 2 or 3). Interestingly, even though 9 and 10 react readily with excess isocyanide, methylene chloride solutions of 9 and 10 are inert towards reaction with CO at 1 atm and r.t. Furthermore, unlike many TMPP complexes, [Rh(TMPP)2(CNBut)2][BF4]2 (9) does not undergo nucleophilic attack by free TMPP to yield dealkylated .._~. -4 156 complexes. Such behavior suggests that the pendent methoxy groups of TMPP are not tightly bonded to the metal center and that the coordination geometry about rhodium is four coordinate. Indeed, this is the case as evidenced by X—ray crystallography . Thus far, only tert-butyl and iso-propyl isocyanide ligands have been to produce stable Rh(II) isocyanide adducts of [Rh(n3-TMPP)2][BF4]2. Reactions of 3 with other isocyanides, such as cyclohexyl or n-butyl isocyanide, result in the initial formation of dark colored solutions that eventually decompose to form a mixture of isocyanide products as evidenced by infrared spectroscopy. The exact nature of these products has not yet been elucidated. However, by analogy with the chemistry of 9 and 10 with excess isocyanide, it is likely that a Rh(II) complex is formed initially, but is subject to facile isocyanide dissociation. The liberated isocyanide in turn reacts with the other Rh(II) isocyanide complexes. Clearly, more evidence is required to support this hypothesis and to formulate the products. More intriguing than the reactions with cyclohexyl and n-butyl isocyanide, however, is the chemistry of [Rh(n 3-TMPP)2][BF4]2 (3) with methyl isocyanide. Unlike the reactions of 3 with longer chain isocyanides in which the products remain in solution, addition of methyl isocyanide to methylene chloride solutions of 3 results in the immediate formation of a red precipitate. This solid is insoluble in most common organic solvents except for acetonitrile, in which the compound exhibits only limited solubility. Infrared spectroscopy reveals a single CEN stretching vibration at 2238 cm1 and a broad band at 1060 cm'1 indicating that complex is cationic. The solid is diamagnetic as evidenced by 1H and 31P spectroscopy. The 1H spectrum of the solid exhibits one triplet resonance representing all meta-protons of the phosphine; the observed coupling pattern for the meta-protons indicates a 157 trans-arrangement of two phosphines about the metal. The 31P NMR spectrum of the solid in CD3CN exhibits two idependent resonances that nearly overlap each other. Each resonance is split into a doublet, presumably due to Rh-P coupling, but there no evidence for the presence of P-P coupling. Such a situation might arise if the complex existed in solution as a 1:1 ratio of two independent geometric conformers in which the trans phosphorus nuclei of each conformer are magnetically equivalent.12 Further work is needed to determine the identity of this product. B. Magnetic and EPR Spectroscopic Properties of [Rh(TMPP)2(CNR)2][BF4]2 (R = But, Pri). The paramagnetism of [Rh(TMPP)2(CNBut)2][BF4]2 (9) was probed by several spectroscopic and magnetic techniques. The 1H and 31P NMR spectra of [Rh(TMPP)2(CNR)2][BF4]2 are broad and essentially featureless, consistent with the formulation of 9 and 10 as paramagnetic species. Solid state and solution magnetic susceptibility studies confirmed the presence of an S = 1/2 ground state for the isocyanide complexes. A solid state magnetic susceptibility measurement of 9 at 299 K led to a ueff value of 2.04 B.M. A diamagnetic correction of -880 x 10'6 cgs was applied based on -24 x 10'6 cgs for Rh2+, -39 x 10-6 cgs for [BF4]': -330.8 x 10-6 cgs for TMPP and -58.5 x 10- 6 cgs for CNBut.13 Variable temperature magnetic susceptibility measurements for 9 were made over the temperature range of 5 - 380 K at a field strength of 500 G; Consistent with a simple paramagnet, the sample displayed Curie-Weiss behavior over this temperature range with 9 = -3.35 K. Solution susceptibility studies by the Evans method yielded a ueff of 2.20 B.M. at 293 K, also consistent with the presence of an S = 1/2 ground state.14 The paramagnetism of the samples was further examined by EPR spectroscopy. The solid-state EPR spectrum of a polycrystalline sample of 9 158 at 100 K (Figure 26) shows an axial signal with g = 2.45 and g1] = 1.96 with hyperfine“ coupling to 103Rh (1:1/2) in the g1] region (A11 = 62 G). The EPR spectrum in a 1:1 Me-THF/CH2C12 glass at 100 K exhibits a signal similar to that observed in the solid state except the signal has become rhombic and the gx and gy regions show hyperfine coupling to 103Rh as well (Figure 27). Although Rh hyperfine coupling in this region is rarely observed, it has been reported by Wilkinson et al. for the organometallic species Rh(2,4,6- Pri3CGH2)2(tht)2.15 [Rh(TMPP)2(CNPri)2][BF4]2 (10) exhibits an EPR spectrum similar to that observed for 9 except that the gx and gy tensors are nearly coincidental as shown in Figure 28. The rhombic signal with gx ~ gy > gz, observed for both 9 and 10, is characteristic of a (1,2 ground state. C. Redox Chemistry of [Rh(TMPP)2(CNR)2] [BF412 (R = But, Pri) A cyclic voltammogram of [Rh(TMPP)2(CNBut)2][BF4]2 in 0.1 M TBABF4-CH2C12 shows a reversible couple at Em: -0.04 V vs Ag/AgCl, corresponding to a one-electron reduction to Rh(I). Not surprisingly, this process is shifted to more positive potentials relative to that of [Rh(n3- TMPP)2][BF4]2 (3), due to the electron-withdrawing effect of the n-acceptor ligands.16 Comp0und 9 can be chemically reduced in the presence of cobaltocene to give the yellow Rh(I) complex [Rh(TMPP)2(CNBut)2][BF4] (11). Examination of the cyclic voltammogram for 11 shows that it is identical to that of 9, except that the redox process at E1/2 = -0.04 V corresponds to an oxidation of the compound. The infrared spectrum of 11 shows a strong band v(CEN) = 2118 cm'l, shifted to lower energy than the corresponding stretch in 9 due to increased n-back-bonding upon reduction from Rh(II) to Rh(I). Although the complex is expected to exhibit a simple trans-nl-phosphine 159 9.1.: 2.45 g" = 1.96 1918 2334 2750 3167 3583 3998 GAUSS Figure26. EPR spectrum of [Rhn(TMPP)2(CNBut)2][BF4]2 (9) in the solid-state at 100 K. um M 2: a as» $536545 . a E :3 43::acsmzeaamzevia as 85.58% mum .5 2.52 Amm3 Oj [PM-l RhPIII] 2+ Me Me —I MeO) ax,eq-[RhmmM -mp-0),]1+ LO(\T‘ (14> [flu RhHI P) NUC: NEE/0’ ;h|)/Me 1+ M Me —] III _ _ - 2+ 0 0 0) ax-[Rh (n3 'I'MPP)(T]3 TMPP 0)] (5) Figure 32. ax.ax-[Rhm(n)3-TI\/IPP-0)]1+ (15) Possible dealkylation products resulting from nucleophilic attack on ax-[RhHI(n3-TMPP)(113-TMPP-0)][BF4]2 (5). 188 detected by 1H NMR spectroscopy. Nucleophilic attack on a coordinated methoxy group is expected to be influenced by a variety of factors including steric hindrance of the ancillary ligands, electrophilicity of the -OCH3 moiety and basicity of the attacking nucleophile. In light of this, it is not surprising that minor variations in reaction conditions lead to different relative amounts of the isomers 14 and 15. This concept is illustrated by the fact that demethylation of 5 with a weaker nucleophile, such as iodide, produces 15 exclusively. The phenoxide group of ax-[Rhm(n3-TMPP)(n3-TMPP-0)][BF4]2 is easily protonated in the presence of excess HBF4. However, the expected product [Rhm(n3-TMPP)(n3-TMPP-0H)][BF4]3 appears to be present in only minor amounts as evidenced by 31P NMR spectroscopy. The 31P NMR spectrum of the major product exhibits a doublet at 5 = + 36.2 ppm (1JRh_p = 129.7 Hz); this is indicative of two chemically equivalent phosphines. One possible explanation is that upon protonation of 5 to form [RhHI(n3- TMPP)(n3-TMPP-0H)][BF4]3, a dealkylation reaction occurs similar to the one that transforms [Rh(n3-TMPP)2][BF4]3 (4) to 5. By analogy with the dealkylation of 4, such a process is not unreasonable, if one considers the high positive charge concentrated on the Rh atom. The absence of anionic ligands, which would tend to offset this charge, creates a highly electrophilic metal center that may activate another coordinated methoxy group. Following demethylation, the newly formed phenoxide ligand is quickly protonated to form [Rh(n3-TMPP-OH)2][BF4]3. A axial orientation of the phenol groups would result in an arrangement where both phosphino-phenol groups are chemically equivalent. Clearly, the preceding arguments represent just one of several possible scenarios and further work is needed to either prove or disprove the hypothesis. 189 D. Chemistry of eq-[RhH(TMPP)(TMPP-0)][BF4] (12) with CO One of our primary goals in preparing a demethylated analog of [Rh(n3-TMPP)2][BF4]2 (3) was to modify the reactivity of the system by reducing the flein'bility of the ligand set. We argued that dealkylation of the pendent methoxy group would result in the formation of a less labile phenoxide interaction, thereby constraining the coordination environment about the metal. Based on the recent fascinating work of Wayland et al. on Rh(II) porphyrins, we anticipated that by introducing a more rigid ligand framework, we may inhibit disproportionation and therefore isolate a Rh(II) carbonyl species. The presence of an anionic phenoxide ligand is also expected to help stabilize the Rh(II) oxidation state with respect to Rh(I) and further deter disproportionation. Moderate purging of a methylene chloride solution of eq- [RhH(TMPP)(TMPP-0)][BF4] (12) with CO results in an immediate change in the solution color from red to dark green. An infrared spectrum measured within the first 10 minutes exhibits a set of moderate intensity bands at 2084, 2068, and 1990 cm'l. Moreover, two weak, higher energy bands appear at 2186 and 2136 cm'l. Upon continued purging of CO, the green color quickly converts to yellow/orange. An infrared spectrum of the solution at this stage reveals an absence of the two higher energy bands and an increase in the intensity of the three lower energy absorptions. Removal of the CO atmosphere by purging with argon causes a reduction in the bands at 2084, 2068, and 1990 cm'l, but the original red solution color never reappears indicating that the overall addition of CO to eq-[RhH(TMPP)(TMPP-0)l[BF4] (12) is not a reversible process. Intrigued by the possibility that the higher energy intermediate bands could be Rh(II) carbonyl adducts, the reaction was monitored by EPR 190 spectroscopy in order to detect the formation of paramagnetic intermediates. A frozen solution of 12 dissolved in a 1:1 mixture of CH2Cl2/Me-THF in an EPR tube was exposed to CO; the sample was gently warmed, and upon reaction was quickly refrozen. The EPR spectrum of the frozen solution at 100 K (Figure 33) revealed the presence of a new paramagnetic species (g1 ~ 2.31, g2 ~ 2.27, g3 ~ 1.98, A3 ~ 14.6 G) in addition to unreacted 12. A similar experiment performed with labeled 13CO, resulted in an EPR spectrum that showed complete conversion to a new Rh(II) species (g1 ~ 2.31, g2 ~ 2.27, g3 ~ . 1.99, A3 ~ 48 G). The lack of significant 13CO coupling to any of the g tensors indicates there is little electron density residing on the CO ligand and that the radical is primarily metal-based. This behavior is in contrast to the significant delocalization of electron density between the Rh and CO moiety observed for the Rh(II)-porphyrin-carbonyl complexes.8 In fact it is exactly the delocalization that induces the carbon coupling reactions in these systems. Unfortunately, the ability of Rh(II)-TMPP complexes to effect similar coupling reactions as those observed for the porphyrin systems appears to be limited. 1 Although EPR spectroscopy provided evidence for the formation of transient Rh(II) CO adducts, it was evident that this was followed by a disproportionation pathway. In order to identify the key participants in this pathway, the progress of the reaction was also monitored by 31P NMR spectroscopy. The 31P NMR spectrum of a solution of eq-[RhH(TMPP)(TMPP- 0)][BF4] (12) in CD30N exposed to CO for one minute reveals that a complex mixture of diamagnetic species is formed. Careful inspection reveals that two of the products are a mixture of ax,eq and ax,ax isomers of [Rhmm3-TMPP- 0)2][BF4], characterized in the previously mentioned dealkylation chemistry. In addition, two unidentified species are also present in solution, a doublet at 191 l l 2552.1 2859.85 3077.52 3285.38 3493.14 3700. Field (GAUSS) Figure 33. EPR spectrum of a frozen 1:1 CH2C12/Me-THF solution of eq- [Rhn(TMPP)(TMPP-0)][BF4] (l2) exposed to CO. 192 5 = + 0.5 ppm (lJRh_p = 132.8 Hz) and an unsymmetrical species characterized by two ABX resonances at 8 = + 40.9 ppm (1JRh_p = 137.3 Hz, 2Jp_p = 16.7 Hz) and 6 = + 24.5 ppm (lJRh_P = 122.7 Hz, 2Jp_p = 16.7 Hz). The presence of [Rh(n3-TMPP-0)2][BF4] implies that certain reactions must participate in the overall reaction pathway. The formation of [Rh(n3-TMPP- 0)2]+ necessitates an initial oxidation of 12 to 13 By analogy to the CO chemistry of [Rh(n3-TMPP)2][BF4]2 (3 ), oxidation of 12 to 13 could, presumably, arise from reaction between 12 and a CO adduct of 12 Such an event is not unreasonable considering the accessible nature of the Rh(II)/Rh(III) redox couple for 12 (El/2(ox) = - 0.02 V). The presence of [Rh(n3- TMPP-0)2]1+ also implies that, during the course of the reaction, TMPP dissociates from one of the products and acts as a nucleophile in the dealkylation of [RhIII(n3-TMPP)(T]3-TMPP-0)]2+. This conclusion is supported by the presence of a significant amount of [TMPP-CH3]+ in the NMR spectrum of the reaction. 4. Discussion The results presented in this chapter demonstrate the susceptibility of RhHJH-TMPP complexes to dealkylation of a coordinated methoxy group. The thermodynamic driving force behind this reaction is undoubtedly the high positive charge on the metal center coupled with activated methyl groups. Although TMPP has been shown to be a highly basic 2 e' donor phosphine ligand with the lone pair, when the ligand is participating in an 113 bonding mode it is a relatively poor 6 e' donor due to the weak donating ability of the pendent ether groups. Consequently, the neutral phosphine is unable to effectively counterbalance the high positive charge on the metal center, and as a result, the coordinated methoxy groups become activated 193 towards dealkylation. As the metal oxidation state increases, and hence the overall charge on the complex, dealkylation becomes increasingly more thermodynamically favorable. This point is nicely demonstrated by a comparison of the relatively stability of the [RhH(n3-TMPP)2]2+ (3) and [RhIH(n3-TMPP)2]3+ (4) complexes towards dealkylation; compound 4 readily demethylates in the presence of very weak nucleophiles such as [BF4]', while 3 generally requires strong nucleophiles. The Rh(III) oxidation state appears to be highly susceptible, as evidenced by the observation that ax-[Rhm(n3- TMPP)(n3-TMPP-0)][BF4J2 will undergo a second dealkylation in the presence of free iodide. The introduction of a phenoxide donor into the metal coordination sphere sets up the possibility of preparing structural isomers based on the relative positions of the phosphine and phenoxide groups. As a result, there are two primary structural conformations, assuming a cis-phosphorus donor atom geometry is maintained; the phenoxide group may either be equatoiral or axial position with respect to the equatorial plane containing the phosphorus atoms. The favored structural conformation is highly dependent on the metal oxidation state. The relationship between the various isomers of the dealkylated and non-dealkylated complexes 3, 4, 5, 12 and l3is depicted in Figure 34. An important underlying principle that is pervasive in the chemistry of these phosphine complexes is the difference in substitutional lability between the +2 and +3 oxidation states of rhodium. The Rh(III) oxidation state, being a (16 metal ion, is relatively substitutionally inert and therefore any structural rearrangements that may occur will be very slow or kinetically inhibited. In contrast, paramagnetic Rh(II) complexes are substitutionally labile and as a result, the distorted RhH-TMPP complexes are less structurally rigid. This lack of structural rigidity is demonstrated by 194 .3 as .2 .m .2. .2 822288 33133852.: was “coca—.3325 me 98:83 2.323 2: cuckoo: 3328528 on» man—33% cannon casuaom .3" 953% as @ s N I a:o-&§-.§%§552§I... .Nxo..:2._.-.:vag-.55§_-s a. a: .5525 02 02 3 o: 2 oz 2 02 02 22 O \ O _ . . _ . _ 02 02 ..r .2 .2 .._| .2 .2 .._I 8&0 583220. .5 .oz 3: 6 15:02.6. ._29:255222255222» 23.52.5ch 22 oz 3 \o \02 02 \0 \oz 220 o flerKou o 5r\ou OZOAAWMI l02\ r\ n—zm— _..¢\\Am j flmm ...Qsa 220 .o u o Q How: /¥O among... .522 moo—.4 /¥O 195 the facile axial to equatorial isomerization of [RhH(TMPP)(TMPP-OlllBF4] (12), while the analogous rearrangement for the Rh(III) d6 complex is exceedingly slow. Curiously, in each of the complexes described in this chapter, the phosphine lie cis to each other rather than in the sterically more logical trans orientation. Evidently, the preference for the cis geometry is electronic and not kinetic in origin. Presumably, the presence of the weak ether donors requires that the phosphine ligands remain cis in order to maximize electron donation to the relatively electron-poor metal center. Several observations support this conclusion. Previously, we noted that [Rh(n3-TMPP)2][BF4]2 (3) was present as a cis-phosphine conformation and that it could not be thermally converted to the trans orientation. One could conceivably argue that the steric bulk of the ligands prevents isomerization to the steric favored conformation and that the cis orientation of the phosphines resulted from the original attack of the phosphine on [Rh2(MeCN)10][BF4]4. This seems unlikely in light of the facile rearrangement of ax-[Rhm(n3-TMPP)(n3-TMPP- 0)]2+ (5) upon reduction to the corresponding d7 complex [RhII(TMPP)(TMPP-0)]1+ (12). The substitutionally labile nature of the d7 complexes allows the phenoxide group to readily change its relationship (axial/equatorial) to the neutral phosphine, yet in compound 3 the phosphorus atoms remain cis to each other. Moreover, introduction of strong donor ligands such as CO and CNR results in an immediate structural rearrangement to a trans phosphine geometry as exhibited by complexes 6 - l 1. This point is dramatically illustrated by the oxidation of [RhI(TMPP)2(CO)]1+ (7) to [RhH(113-TMPP)2]2+ (3). Although the phosphines are trans in the initial Rh(I)CO complex, oxidation followed by CO dissociation results in phosphine isomerization to the thermodynamically 196 favored cis orientation. These observations clearly support that the disposition of the phosphine ligands in these Rh(II) and Rh(III) complexes is not kinetic in origin but instead thermodynamic. A final point regarding the dealkylation of TMPP to form the anionic phosphino-phenoxide needs to be addressed, namely, the effect that dealkylation has on the chemistry of these complexes. An important consequence of demethylation is that it limits the flexibility of the ligand by anchoring it at two positions rather than only one. Furthermore, the formation of a relatively non-labile metal-phenoxide bonds reduces the number of potential coordination sites for an incoming substrate. Nonetheless, these phenoxide groups do not occupy all of the coordination sites. The presence of two or three labile metal ether interactions still leaves open the possibility for further chemistry. Ironically, the formation of phenoxide donors may potentially induce higher lability in the remaining coordinated methoxy groups by providing more electron density at the metal. This in effect will reduce the electrophilic character of the metal and decrease the need for electron donation by the pendent ether groups and subsequently making them more labile. This will be particularly true for higher valent metal centers. Consequently, [RhIH(T]3-TMPP)(T]3-TMPP-0)]2+ (5) and [Rh111(n3-TMPP-0)2]1+ (14, 15) may prove to be more reactive towards addition of small substrates to the metal center. Comprehensive studies of these systems are needed to establish the use of phosphino-alkoxide groups as ancillary ligands for late transition metal complexes. 197 List of References 1. (a) Mehrotra, R. C.; Agarwal, S. K.; Singh, Y. P. Coord. Chem. Rev. 1985, 68, 101. (b) Bryndza, H. E.; Tam, W. Chem. Rev. 1988,88, 1163. 2. (a) Bryndza, H. E.; Calabrese, J. C.; Marsi, M.; Roe, D. C.; Tam, W.; Bercaw, J. E. J. Am. Chem. Soc. 1986, 108, 4805. (b) Bryndza, H. E.; Fong, L. K.; Paciello, R. A.; Tam, W.; Bercaw, J. E. J. Am. Chem. Soc. 1987, 109, 1444. (c) Bryndza, H. E.; Domeille, P. J.; Tam. W.; Fong, L. K.; Paciello, R. A.; Bercaw, J. E. Polyhedron 1988, 7, 1441. (d) Hartwig, J. F.; Andersen, R. A.; Bergman, R. G. Organometallics 1991, 10, 1875. 3. For leading references see (a) Loren. S. D.; Campion, B. K.; Heyn, R. W.; 'I‘illey, T. D.; Bursten, B. E.; Luth, K. W. J. Am. Chem. Soc. 1989, 111, 4712 and references therein. (b) Glueck, D. S.; Newman-Winslow, L. J .; Bergman, R. G. Organometallics 1991, 10, 1462. (c) Hartwig, J. F.; Andersen, R. A.; Bergman, R. G. J. Organomet. Chem. 1990, 394, 417. (d) Kegley, S. E.; Schaverian, C. J.; Freudenberger, J. H.; Bergman, R. G. J. Am. Chem. Soc. 1987, 109, 6563. 4. (a) Empsall, H. D.; Hyde, E. M.; Jones, C. E.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1974, 1980. (b) Mason, R.; Thomas, K. M.; Empsall, H. D.; Fletcher, S. R.; Heys, P. N.; Hyde, E. M. Jones, C. E; Shaw, B. L. J. Chem. Soc., Chem. Comm. 1974, 612. (c) Empsall, H. D.; Hyde, E. M.; Shaw, R. L. J. Chem. Soc., Dalton Trans. 1975, 1690. (d) Empsall, H. D.; Heys, P. N.; McDonald, W. 8.; Norton, M. C.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1978, 1119. 5. (a) Dunbar, K. R.; Quillevéré, A. Inorg. Chem. submitted for publication. (b) Quillevéré, A. Ph. D. Dissertation, Michigan State University, 1992. 6. (a) Chen, S. J .; Dunbar, K. R. Inorg. Chem. 1991, 30, 2018. (b) Chen, S. J.; Dunbar, K. R. Inorg. Chem. 1990, 29, 588. (c) Dunbar, K. R.; Saharan, V.; Matonic, J. M. manuscript in preparation. 7. Gray, H. B.; Hendrickson, D. N.; Sohn, Y. S. Inorg. Chem. 197], 10, 1559. 8. (a) Wayland, B. B.; Sherry, A. E. J. Am. Chem. Soc. 1989, 111, 5010. (b) Wayland, B. B.; Sherry, A. E.; Poszmik, G.; Bunn, A. G. J. Am. Chem. Soc. 1992, 114, 1673. CHAPTER VIII CHEMISTRY OF TRIS(2,4,6-TRIlVIETHOXYPHENYL)PHOSPHINE WITH RHODIUM (I) AND IRIDIUM (I) OLEFIN COMPLEXES 198 199 1. Introduction In chapter III, we demonstrated that it was possible to prepare stable Rh(II) TMPP complexes from the solvated dinuclear precursor [Rh2(MeCN)10][BF4]4. In spite of this success, we were interested in synthesizing complexes of the general formula [Rh(TMPP)2]n+ (n: 1, 2, 3) from more conventional starting materials. The development of such synthetic strategies is particularly important for the isolation of homoleptic Ir/TMPP complexes, since the analogous solvated dinuclear Ir complex is unknown. Others have established that Rh(II) and Ir(II) phosphine complexes can be prepared by reaction of M(III) trihalides with excess phosphine in alcohol.1 These conditions, however, are unsuitable for the preparation of [M(TMPP)2]n+ (n: 1, 2, 3), because of the susceptibility of these species to dealkylation. An alternative approach is to begin with partially solvated Rh(I) and Ir(I) olefin complexes which are prepared in situ by halide abstraction from the corresponding halide bridged dinuclear metal olefin complexes.28L Such systems have proven to be good precursors for a variety of group 8 metal phosphine complexes.2 We set out to adapt this methodology to the preparation of Rh(I) and Ir(I) complexes of type [MI(TMPP)2]1+. This chapter describes our efforts to prepare homoleptic TMPP complexes using Rh(I) and Ir(I) olefin complexes. In addition, the synthesis and characterization of the iridium dicarbonyl complex [Ir(TMPP)2(CO)2][BF4] is presented. 2. Experimental A. Synthesis All reactions were carried out under an argon atmosphere by the use of standard Schlenk-line techniques unless otherwise stated. TMPP (1) was 200 prepared as described in chapter II. The starting material, [Ir(cod)Cl]2, was prepared from IrCl3 - xHZO using a standard literature procedure.3 [Rh(CZH4)2Cl]2 was prepared according to a literature procedure.4 [Rh(cod)Cl]2 (Strem) and NOBF4 (Strem) were obtained from commercial sources and used as received. The oxidizing agent, [(p-BrCGH4)N][BF4], was prepared by oxidation of the amine, (p-BrC6H4)N, with I2 in the presence of Ag“ following a modification of the procedure described by Bell et al.5 (1) Preparation of [Rh(cod)(n2-TMPP)I[BF4] (16) To a mixture of [Rh(cod)Cl]2 (0.200 g, 0.406 mmol) and AgBF4 (0.158 g, 0.812 mmol) was added 5 mL of THF. The resulting yellow suspension was stirred for 5 min before being filtered through a Celite plug into a 100 mL 3- necked flask equipped with an addition funnel containing a solution of TMPP (0.432 g, 0.811 mmol) in 15 mL of THF. The yellow solution was then cooled to approximately - 40°C with a dry ice / MeCN slush. The THF solution of TMPP was carefully added dropwise to the chilled solution over a period of 30 min. The reaction was stirred for an additional hour at -40°C, during which time a yellow solid separated from solution. Additional product was obtained from the solution by slow addition of diethyl ether (50 mL) while the solution was being stirred. The cold suspension was filtered in air and the yellow solid was washed with several portions of diethyl ether (4 x 10 mL) and dried under reduced pressure for 1-2 h; yield 0.574 g (85%). Anal. Calcd for C35H4509F4BPRh: C, 50.62; H, 5.46. Found: C, 50.16; H, 5.58. 1H NMR (CD2C12) 8, ppm: 3.68 (s, 36H, o-OCH3), 3.84 (s, 18H, p-OCH3), 6.14 (d, 4Jp_H = 3.6 Hz, 12H, m-H), 1.80 (br, cod), 2.36 (br, cod). 31P NMR (Cchlzl 8, ppm: + 1.4 (d, 1JRh-p = 137.3 Hz). Cyclic voltammogram (0.2 M TBABF4 / CHzClg, vs Ag/AgCl): E1(p,a) = + 0.60 V, E2(p,a) = + 1.16 V, Ep,c = + 0.92 V. Mass spectrum (FAB, 3-nitrobenzyl alcohol) m/z: 742 (Rh(cod)(n2-TMPP)+). 201 (2) Preparation of Rh(cod)(n2-TMPP-0) (17) A solution of [Rh(cod)(n2-TMPP)][BF4] (16) (0.200 g, 0.240 mmol) and TMPP (0.128 g, 0.240 mmol) in 5 mL of MeCN was stirred at r. t. for 12 h. The solvent was removed under vacuum to yield a pale yellow residue. The yellow product was extracted by addition of 10 mL of Et20 and a minimal amount of THF (an amount sufficient to dissolve the yellow solid completely). The solution was filtered to remove the [TMPP-CH3][BF4] by-product. The solvent was then removed under vacuum to produce a yellow powder; yield, ' 0.142 g (81%). 1H NMR (CD2C12) 8, ppm: -OCH3, 3.35 (s, 3H), 3.55 (s, 12H), 3.68 (s, 3H), 3.80 (s, 6H); m-H, 5.45 (t, 1H), 5.72 (dd, 1H), 6.04 (d, 4Jp_H = 3.3 Hz, 4H); cod, 1.86 (br), 2.38 (br), 3.38 (br), 4.91 (br). 31P NMR (CDzClz) 8, ppm: + 8.0 (d, lJRh-P = 157.2 Hz). Mass spectrum (FAB, ONPO) m/z: 728 (Rh(cod)(n2-TMPP-0)+). (3) Reaction of [Rh(C2H4)2Cl]2 with TMPP A solution of TMPP (0.548 g, 1.03 mmol) in 10 mL of MeCN was added dropwise to a cold suspension (0°C) of [Rh(C2H4)2Cl]2 (0.100 g, 0.257 mmol) and KBF4 (0.064 g, 0.514 mmol). After the addition of the phosphine was complete, a slight vacuum was applied periodically to induce loss of ethylene. The reaction was allowed to warm slowly to r. t. After 2 h of stirring, the solvent was removed under reduced pressure. The resulting residue was dissolved in CH2012 (10 mL) and filtered through a Celite plug to remove undissolved KCl. The filtrate was evaporated to a solid under reduced pressure which was washed with diethyl ether (2 x 10 mL) and dried in vacuo. 1H and 31P NMR spectroscopy showed that the solid was comprised of a mixture of [TMPP-CH3]+ and [TMPP-CH2C1]+ together with a Rh species similar to the species observed upon reduction of [Rh(n3-TMPP)2][BF4]2 (3) with cobaltocene. 202 (4) Preparation of Ir(cod)(n2-TMPP-0) ( 18) To a solution of [Ir(cod)Cl]2 (0.150 g, 0.223 mmol) and KBF4 (0.056 g, 0.446 mol) at 0°C was added dropwise to a solution of TMPP (0.475 g, 0.892 mmol) in 10 mL of MeCN. The solution color gradually converted from yellow to orange as the reaction was allowed to warm to r. t. After stirring for 2 h, the solvent was removed under vacuum to yield an orange residue. The solid was taken up with 10 mL of CH2Cl2 and filtered through a Celite plug to remove the KC] by-product. The Celite plug was further washed with 5 mL of CH2012 to ensure complete transfer of the product. The volume of the solution was reduced to ~ 5 mL, and diethyl ether (20 mL) was slowly added while the solution was being stirred. The resulting yellow/white precipitate was removed by filtration and the orange filtrate was concentrated to approximately 5 mL. To the orange solution was added 20 mL of hexanes. The volume was again concentrated to 3-5 mL resulting in the formation an orange precipitate. The solid was filtered in air, washed with hexanes (2 x 10 mL), and dried under reduced pressure; yield, 0.302 g (83%). 1H NMR (CD3CN) 8, ppm: -OCH3, 3.36 (s, 3H), 3.49 (s, 12H), 3.68 (s, 3H), 3.79 (s, 6H); m-H, 5.55 (t, 1H), 5.77 (dd, 1H), 6.09 (d, 4JP-H = 3.6 Hz, 4H). (5) Preparation of [Ir(TMPP)2(CO)2] [BF4] (19) A 100 mL 3-necked flask, equipped with an addition funnel, gas inlet and a septum, was charged with [Ir(cod)Cl]2 (0.200 g 0.30 mmol), NaBF4 (0.065 g, 0.60 mmol) and MeCN (5 mL). The resulting solution was cooled to 0°C and gently purged with CO. A solution of TMPP (0.634 g, 1.20 mmol) in 10 mL of MeCN was added dropwise to the CO saturated solution over a period of ten minutes. After the addition was complete, the CO purge was discontinued, and the reaction was stirred under a CO atmosphere at 0°C. After 2 h, the solvent was removed under vacuum. The resulting residue was 203 dissolved in 5 mL of CH2C12 and filtered through a Celite plug. The Celite was further washed with 5 mL of CHzClZ. The orange solution was concentrated to 4-6 mL and 35 mL of Et20 was added slowly. The resulting bright orange precipitate was filtered in air; washed with 4 x 10 mL of diethyl ether and dried in vacuo; yield 0.666 g (80%). Anal. Calcd for C57H68020F4BC12P2Rh: C, 46.10; H, 4.62. Found: C, 46.75; H, 4.88. IR (CH2C12) cm'1: v(CO), 1993 (vs), 1946 (w). IR (Nujol, CsI) cm'l: v(CO), 1987 (s), 1942 (m). 1H NMR (CD3CN) 8, ppm: 3.38 (s, 36H, o-OCH3), 3.77 (s, 18H, p-OCH3), 6.06 (t, 4Jp_H = 2.0 Hz, 12H, m-H). 31P NMR (CD013) 8, ppm: - 39.0 (s). Electronic absorption spectrum (CH2C12) kmax, nm (e): 547 (449), 490 (2300), 455 (sh) 386 (4510) 315 (sh). Cyclic voltammogram (0.1 M TBABF4 / CH2C12, vs Ag/AgCl): Ep,a = + 0.82 V, E1(p,c) = + 0.46 V, E2(p,c) = - 0.81 V. (6) Reaction of TMPP with [Ir(TMPP)2(CO)2] [BF4] (19) A quantity of [Ir(TMPP)2(CO)2][BF4] (19) (0.100 g, 0.071 mmol) and TMPP (0.038 g, 0.071 mmol) was dissolved in 5 mL of MeCN. The orange solution was stirred at r. t. for 12 h. An infrared spectrum of the solution revealed that no reaction had occurred during this period. A condenser was connected to the flask and the solution was refluxed for 24 h during which time the solution color turned from orange to yellow. An IR spectrum of the solution was obtained after this period of time: IR (CH3CN) cm‘1: v(CO), 2051 (s), 1973 (s),1899 (w). The reaction was refluxed for an additional 24 h, after which time a small aliquot was removed and an IR spectrum was measured: IR (CH3CN) cm'1: v(CO), 1998 (w, br) 2018 (w, br). The solvent was then removed under vacuum and the solid was redissolved in a 2:1 mixture of THF and diethyl ether. The yellow solution was filtered and pumped to dryness. The resulting solid was washed with diethyl ether (2 x 5 mL) and dried in 204 vacuo. The 1H NMR spectrum of the solid in CD3CN revealed the presence of [TMPP-CH3]+, but no other resonances were easily discernible. (7) Oxidation of [Ir(TMPP)2(CO)2][BF4] (19) (i) Chemical oxidation with NOBF4. In a typical reaction, a solution of [Ir(TMPP)2(CO)2][BF4l (19) (0.100 g, 0.071 mmol) and NOBF4 (0.009 g, 0.0.71 mmol) in 5 mL of MeCN was stirred at r. t. In the first minute a vacuum was applied to remove NO(g) or CO that may evolve from the reaction. Within several minutes, the solution color changed from orange to orange-red. An' aliquot was removed after 30 min and its IR spectrum was recorded. IR (CH3CN) cm'1: v(CO), 2079 (s), 2109 (w), 2148 (w), 2029 (vw). The reaction was stirred overnight, after which time the solution was pump to dryness, washed with 10 mL of Et20 and dried in vacuo; yield, 0.053 g of crude product. (ii) Chemical Oxidation with [(p-BngH4)3N][BF4]. In a typical reaction, a mixture of [Ir(TMPP)2(CO)2][BF4] (19) (0.100 g, 0.071 mmol) and [(p-BrC6H4)3N"_I[BF4] (0.041 g, 0.071 mmol) was dissolved in 5 mL of MeCN. The solution immediately became dark red/orange in color. The reaction was periodically subjected to a vacuum to help remove evolved CO from the solution. Within 30 min, the solution color began to lighten and an IR spectrum of the solution showed the presence of unreacted 19 together with several other v(CO) bands indicative of a new compound or compounds; IR (CH3CN) cm'1: v(CO), 2079 (s), 2110 (m), 2148 (m); 19, 1991 (s), 1942 (w). The reaction was stirred overnight during which time the solution color became pale green and a pale precipitate had settled out. An aliquot of the green solution was removed and its IR spectrum was recorded; IR (CH3CN) cm'l: v(CO), 2079 (vs), 2027 (w), 2148 (vw), 2110 (vw), 2046 (vw), 2009 (w). The solution was filtered and evaporated to a residue under reduced pressure 205 to yield a pale yellow/green solid. 1H NMR (CD3CN) 8, ppm: -OCH3, 3.46 (s), 3.82 (s); m-H, 6.20 (t, 4JP-H = 2.2 Hz). (iii) Bulk Electrolysis. Bulk electrolysis of [Ir(TMPP)2(CO)2][BF4] (19) was performed in a four-compartment electrolysis cell separated by coarse porosity sintered glass frits (Figure 35). A quantity of [Ir(TMPP)2(CO)2][BF4] (19) (0.050 g, 0.036 mmol) was added to a degassed CH2C12 solution of 0.05 M [(Bun)4N][BF4_'| in the working compartment of the cell. The solution was electrolyzed at a potential of E = + 1.0 V using a Pt gauze worh'ng electrode. After 1 h, the solution color had changed from orange to yellow. A small aliquot of the solution was removed and its IR spectrum was recorded. IR (CH2Clz) cm'lz v(CO), 2075 (m). The solution was then transferred via syringe into a Schlenk flask under Ar and was carefully layered with Et20 (5 mL). After diffusion was complete, the remaining yellow solution was decanted away from the [(Bun)4N][BF4] salt that had precipitated. The solvent was removed under vacuum to give a pale yellow solid. 1H NMR (CD3CN) 8, ppm: -OCH3, 3.57 (s), 3.86 (s); m-H, 6.19 (t, 4Jp_H = 1.8 Hz). (8) Reaction of [Ir(TMPP)2(CO)2][BF4] (19) with Iodine An amount of [Ir(TMPP)2(CO)2][BF4l (19) (0.050 g, 0.036 mmol) in 5 mL of MeCN was stirred with one equivalent of 12 (0.009 g, 0.036 mmol). and NaBF4 (0.004 g, 0.036). After stirring for 3 h, the solution color became yellow. A small aliquot was removed and its IR spectrum was recorded: IR (CH3CN) cm'1:v(CO), 2148 (vw), 2073 (m), 2055 (s). The solution was pumped to dryness and dissolved in 5 mL of CHZCIZ. The yellow solution was then filtered through a Celite plug to remove the remaining undissolved sodium salt. The filtrate was evaporated under reduced pressure and dried under vacuum. ? Pt gauze Ag/AgC] electrode reference Pt wire electrode iléi th. fnt L Figure 35. Schematic drawing of the electrolysis cell used in the electrochemical oxidation of [IrI(TMPP)2(CO)2]1+ (19). 207 B. X-ray Crystallography The structure of [Ir(TMPP)2(CO)2][BF4] (19) ° CH2012 was determined by application of general procedures that have been fully described elsewhere.6 Geometric and intensity data were collected on a Rigaku AFCGS diffractometer with graphite-monochromated MoKa (15 = 0.71069 A) radiation and were corrected for Lorentz and polarization effects. Relevent crystallographic parameters for 19 are summarized in Table 15. All calculations were performed with the use of VAX computers on a cluster network within the Department of Chemistry at Michigan State University using the Texsan software package of the Molecular Structure Corporation.7 (1) [Ir(TMPP)2(CO)2][BF4] (19) . CH2C12 (i) Data Collection and Reduction. Single crystals of 19 suitable for X- ray analysis were obtained as a CH2C12 solvate from careful layering of Et20 on a CH2C12 solution of 19. A regular block shaped crystal with approximate dimensions, 0.23 x 0.34 x 0.15 mm3, was selected and secured onto the tip of a glass fiber with epoxy cement. Least-squares refinement of 24 orientation reflections in the range 20 < 20 < 33° resulted in cell constants consistent with a triclinic cell. Intensity measurements were performed at 23 i 3°C using the co-20 scan technique. Reflections with I < 106(1) were re-scanned a maximum of two re-scans and the counts were accumulated to assure good counting statistics. Routine measurement of three check reflections at regular intervals throughout data collection revealed that the crystal had experienced 57% loss in diffraction intensity. A linear decay correction was applied to compensate for the observed loss. In addition, an empirical absorption correction was applied based on azimuthal scans of 3 reflections with Eulerian angle x near 90° resulting in maximum and 208 Table 15. Summary of crystallographic data for [IrI(TMPP)2(CO)2][BF4] - CH2012(19) Formula IrP2020C57H63BF4C12 Formula weight 1592.32 Space group P-1 a, A 13.512(2) b, A 18.348(3) c, A. 13.358(2) or, deg 9726(1) [3, deg 9055(1) 7, deg 95.o2(1) V, A3 3272(2) Z 2 dcalc, g/cm3 1.616 I1 (Mo Ka), cm'1 41.73 Temperature, °C 23i2 °C Ra 0.051 wa 0.065 Quality-of-fit indicatorc 2.19 aR=£ I lFol- IFcIl/ztrol bRw = [2w< I F0 I— I re I)2/zw IF0 |211/2; w = 1/02( IFo I) CQuality-of-fit = [zw( IFo I - IFc [)2/(Nobs-Nparameters)]1/2 209 minimum transmission factors of 1.00 and 0.79. A total of 10,134 data were collected in the range of 4 S 20 S 47°. After averaging equivalent reflections (Rmerge = 3.4%), there remained 9652 unique data of which 6165 had F02 2 36(Fo)2. (ii) Structure Solution and Refinement. The position of the Ir atom was located directly from a Patterson Fourier synthesis. The remaining non- hydrogen atoms were located by application of the program DIRDIF followed by several alternating least-squares cycles and Fourier maps.8 Due to a disorder resulting from several random orientations, the [BF4]' anion was modeled as an ideal tetrahedron with fixed bond distance and angles. With the exception of the [BF4]‘ anion, all non-hydrogen atoms were refined with anisotropic thermal parameters. Hydrogen atoms were included in the structure factor calculation as fixed contributors at calculated positions and were not refined. In the end, final least-squares refinement of 731 parameters gave residuals of R = 0.051 and RW = 0.065 and a quality-of-fit index of 2.19. Not unexpectedly, the highest remaining peak in the final Fourier difference map was 1.19 e'/A3 and was associated with the disordered [BF4]' anion. 3. Results A. Synthesis and Characterization of [M(cod)(n2-TMPP)]1+ (M = Rh) and M(cod)(n2-TMPP-0) (M = Rh, Ir) One equivalent of TMPP readily reacts with the partially solvated Rh olefin complex [Rh(cod)(solvent)2]+, which is formed in situ by halide abstraction using Ag+, to produce [Rh(cod)(n2-TMPP)][BF4] (16) (eq 20). The [BF4]' salt of the cation is relatively insoluble in THF, and as a result, the -_——-1 210 M e1+ e I /Rh [Cl R/\ A IFHFI TMPP m0 RhClR h/ /Rh‘TH (20) main r’ArO Me Ar product is isolated as a fine yellow precipitate in high yield from THF. Although it is anticipated that the TMPP ligand participates in an 112- bonding mode to the metal, thereby filling the vacant fourth coordination site, the 1H NMR spectrum of 16 in CDZCIZ exhibits a highly symmetrical pattern of resonances in which each phenyl ring is equivalent. Clearly, the phosphine is participating in a dynamic process that rapidly exchanges each of the non- coordinated phenyl rings with the coordinated ring. A similar fluxional process was also noted for (n 3 -TMPP)Mo(CO)3 and [Rh(n 2- TMPP)(TMPP)2(CO)][BF4] (7).9 Furthermore, the coordinated diene is also apparently participating in a dynamic process as evidenced by the broad appearance of the 1H NMR resonances for the cod ligand. The 31P NMR spectrum exhibits the expected doublet at 8 = + 1.4 ppm ((1, 1J Rh_p = 137.3 Hz) as a result of one phosphorus nucleus experiencing 103Rh-31P coupling. In an attempt to displace the coordinated olefin to form the homoleptic Rh(I) TMPP complex analogous to 3, [Rh(cod)(n2-TMPP)][BF4] (16) was deliberately reacted with a second equivalent of TMPP. However, instead of substituting the diene, the phosphine behaved as a nucleophile resulting in the dealkylation of a coordinated methoxy group to give [TMPP-CH3]+ (eq 21) and the dealkylated complex, Rh(cod)(n2-TMPP-0) (17), which was isolated from the phosphonium salt by extraction with a mixture of diethyl ether and a minimal amount of THF. Quite unexpectedly, 17 is relatively soluble in ..- 211 Me TMPP M M9 /\ [53 /\ ,0 o Rh Rh n (21) \ \ \/ B, \/ R, 0‘ "’1Ar TMPP CH3 * r" . Me Ar .AI‘Ar Ar' diethyl ether; this is presumably a consequence of its charge neutrality. Crystalline solids may be obtained by careful layering of hexanes onto CH2012 solutions of 17. An infrared spectrum of Rh(cod)(n2-TMPP-O) (17) shows the characteristic bands for coordinated TMPP and confirms the absence of the [BF4]' anion consistent with the neutral formulation of 17. A FAB mass spectrum of the compound exhibits a parent ion peak at 728 m/z in accord with the empirical formula RhPC3409H42. The 1H NMR spectrum of RhI(cod)(n2-TMPP-O) (1 7) in CD2 C12 reveals a relatively symmetric resonance pattern. As was observed for the parent compound, the cyclooctadiene resonances are broad, indicating that a dynamic process is occurring in solution. The most striking feature of the spectrum is the appearance of the resonances corresponding to meta-protons on the phenyl rings of TMPP. The two multiplets that appear at 8 = 5.45 and 5.7 2 ppm are assigned to the proximal and distal meta-protons of the chelating phenoxide ring. The observed coupling pattern arises from coupling of these magnetically distinct resonances to the phosphorus nucleus and to each other. The meta-protons of the two remaining phenyl rings are magnetically equivalent and appear as doublets at 8 = 6.04 ppm (4Jp_H = 3.3 Hz). Evidently, there is still free rotation about the P-C bonds of the non- 212 coordinated phenyl rings, inspite of the fact that one phenyl ring is bonded to the metal. Of further interest is the 31P NMR spectral properties of 17 in CD2C12; the 31P NMR spectrum exhibits a doublet at 8 = - 8.0 ppm (1JRh-p = 157.2 Hz), which is shifted upfield relative to 16. Typically the formation of a five membered chelate ring involving phosphorus results in a significant downfield shift of the phosphorus resonance.10 Apparently, however, in this case the chelation affect is more than compensated for by the formation ofa more strongly donating phenoxide interaction, which increases the electron density at the metal center, thereby further shielding the phosphorus nucleus. The net effect is an upfield shift of the phosphorus resonance. The analogous iridium complex, Ir(cod)(n2-TMPP-O) (18) was prepared by reaction of [Ir(cod)Cl]2 with excess TMPP in the presence of KBF4. Although it was not isolated, [Ir(cod)(n2-TMPP)]+ is believed to be formed in the initial stages of the reaction. This species then undergoes dealkylation in the presence of additional phosphine to yield 18. COmpound l8 exhibits essentially the same solubility and stability properties as the rhodium complex Rh(cod)(n2-TMPP-0) (17). The 1H NMR spectrum of Ir(cod)(n2- TMPP-O) (18) in CD3CN exhibits three resonances at 8 = 5.55 ppm (t), 5.77 ppm (dd), and 6.09 ppm (d, 4Jp_H = 3.6 Hz) in a ratio of 1 : 1 : 4 corresponding to the meta-protons of the phenyl rings. Just as was observed for 17, the spectrum indicates that one ring is tightly bonded to the metal leading to the observation of proximal and distal resonances for the two meta-protons on that ring. The remaining two phenyl rings, however, freely rotate about the P-C bond. 213 B. Synthesis and Spectroscopic Characterization of [Ir(TMPP)2(CO)2l[BF4] (l9) [Ir(TMPP)2(CO)2][BF 4] (19) was synthesized in manner similar to the method used for [Rh(TMPP)2(CO)]+ (eq 22). As in the rhodium chemistry, coordinated diene may be displaced from the iridium complex by CO in the presence of TMPP to give [Ir(TMPP)2(CO)2]+ (19). A solution of 4 equivalents of TMPP in MeCN was added dropwise to a MeCN solution of [Ir(cod)Cl]2 in the presence of MBF4 (M = Ag+, Na+, or K+). The addition was performed under an atmosphere of CO at 0°C with yields typically in the range 80-85%. A drastic reduction in yield was noted when THF was Me- 03:: 1 1+ F‘Ar' /1’Cl1/\\)L<;/1NCM: TMPP Me-Ooc-liar-CO (22) \/Ir\ CllI ry) MeCN \NCMe CO I Ar' ~ 1:‘Ar' A1" used as a solvent instead of MeCN, unlike the synthesis of [Rh(TMPP)2CO]1+ (7 ) which works very well with THF as a solvent. The choice of metathesis reagent apparently does not effect the yield, as Ag+, Na+, and K+ have produced comparable yields f0 19. If the reaction is not purged with CO, the neutral complex Ir(cod)(TMPP-0) (18) is formed instead of 19 As was observed in the analogous Rh system, CO is required to displace the cod. Otherwise, the second equivalent of TMPP acts as a nucleophile and dealkylates the coordinated phosphine instead of displacing the diene. However, unlike [Rh(TMPP)2(CO)2]1+ (6), CO does not reversibly dissociate from 19 to form an 214 iridium monocarbonyl species. Only under more forcing thermal and photolytic conditions is the loss of CO apparent. Unfortunately, such conditions lead to a mixture of intractable products. The infrared spectrum of [Ir(TMPP)2(CO)2][BF4] (1 9) in CH2 C12 exhibits two carbonyl stretching vibrations at v(CO) = 1993 cm’1 vs and 1946 cm'1 (w). In the solid state these bands shift slightly and change intensity to v(CO) = 1987 cm'1 (s) and 1942 cm'1 (In). The 1H NMR spectrum of 191s consistent with a trans disposition of two magnetically equivalent phosphine ligands with resonances appearing at 8 = 3.38 (s, 36 H, o-OMe), 3.77 (s, 18 H, p-OMe), 6.06 (t, 4Jp-H=2.0 hz, 12H, m-H). The 31P NMR spectrum of 19 (CDCl3) exhibits a single resonance for both phosphines at 8 = -39 ppm. The cyclic voltammogram of [Ir(TMPP)2(CO)2][BF4] in 0.1 M TBABF4/CH2C12 shows an irreversible oxidation at Ema: +0.82 V. Coupled with this oxidation are two chemical waves at E = +0.46 V and -0.81 V. Of further interest is the observation that [Ir(TMPP)2(CO)2][BF4] (19) is emissive in the solid state at room temperature. Exposure of solid samples of 19 to long wave ultraviolet radiation produces a brilliant orange luminescence. Emission spectra have been previously found for other square planar Rh(I) and Ir(I) complexes.11 Typically, the emissive state results from MLCT from the metal dz2 orbital to a low-lying ligand orbital of 1: symmetry. When the ligands are n-acceptors, such as CO or CN', the MLCT is interpreted as excitation from the metal dz2 orbital to the 1t* orbital of the ligand.11g By analogy, we expect a similar charge transfer process to be responsible for the emissive properties of 19. In contrast to the solid state behavior, solutions of [Ir(TMPP)2(CO)2][BF4] (19) do not appear to be luminescent at room temperature. This is in agreement with other square planar (18 complexes. 215 C. Crystal Structure of [Ir(TMPP)2(CO)2][BF4] (19) The identity of the product as [Ir(TMPP)2(CO)2][BF4] was confirmed by X—ray crystallography. An ORTEP diagram of the molecular cation is presented in Figure 36. A listing of pertinent bond distance and angles are found in Table 16. As expected, the complex is square planar with the phosphines situated trans to each other. Not surprisingly, the phosphine ligands are monodentate and there are no interacting methoxy groups along the axial direction (r(Ir-O) > 3.0 A). Although the phosphines are ‘ magnetically equivalent in solution, they are not crystallographically identical, as the iridium atom does not lie on a crystallographic symmetry element. The Ir-P distances, 2.338(3) A and 2.345(3) A, are only slightly longer than those in the analogous Rh complex [Rh(TMPP)2(CO)2][BF4] (6) (2.327(1) A and 2.332(1) A). There are no distortions evident in the structure, as the angles between the ligands are nearly ideal. A packing diagram shows that the square planar Ir cations stack along the z-axis, but the closest distance between iridium centers is over 10 A. So it appears that the observed emission properties are associated with the ligands and do not arise from association of cations in the solid state. D. Oxidation of [Ir(TMPP)2(CO)2] [BF4] Bulk electrolysis of [Ir(TMPP)2(CO)2][BF4] in 0.05 M TBABF4 / CH2C12 at a potential of +1.1 V for 1 hour gave a yellow solution that exhibited a carbonyl stretch at v(CO) = 2075 cm'1(m). The 1H NMR spectrum of the solid in CD2C12 showed resonances due to [(nBu)4N]+ in addition to a new TMPP containing species: 8, ppm; 3.57 (s, broad), 3.86 (s), 6.19 (t). The triplet at 6.19 is indicative of a trans diphosphine complex and the breadth of the methoxy resonances suggest that some fluxionality is occurring. Chemical oxidation of [Ir(TMPP)2(CO)2][BF4] with NOBF4 in MeCN 216 C52 C35 n C C48 011 [:30 Q/C'5>3 017 Figure36. ORTEP representation of the molecular cation [IrI(TMPP)2(CO)2]1+ (19). 217 Table lSelected bond distances (A) and angles (deg) for [IrI(TMPP)2(CO)2][BF4] 0 CH2012(19). Atom 1 Atom 2 bond distance Ir(l) P(1) 2.338(3) Ir(l) P(2) 2.345(3) Ir(l) C(55) 1.94(1) Ir(l) C(56) 1.89(1) P(1) C(l) 1.87(1) P(1) C(10) 184(1) P(1) C(19) 1.84(1) Atom 1 Atom 2 Atom 3 bond angles P(1) Ir(l) P(2) 178.4(1) P(1) Ir(l) C(55) 90.0(4) P( 1) Ir(l) C(56) 90.4(4) P(2) Ir(l) C(55) 88.4(4) P(2) Ir(l) C(56) 91.2(4) C(55) Ir(l) C(56) 177.8(6) Ir(1) P(1) C(l) 100.5(4) Ir(l) P( 1) C(10) 115.8(4) Ir(l) P( 1) C(19) 118.8(4) 218 produced a red-orange solution. The infrared spectrum of the solution showed several bands at energies greater than 2000 cm'l; these occur at 2079 cm‘1 (s), 2148 cm‘1 (w), 2109 cm‘1 (w), 2029 cm'1 (vw). Presumably, the intense band at 2079 cm'1 corresponds to the same product formed in the bulk electrolysis experiment and is shifted due to solvent effects (MeCN versus CHzClz). Oxidation of [Ir(TMPP)2(CO)2][BF4] (19) with the amminium salt [(p- BrC6H4)3N][BF4J initially gave a deep blue solution that immediately became. red-orange as the amminium salt was consumed. As in the other oxidation reactions of 19, a vacuum was applied to help remove evolved CO from the solution. After a day of continual stirring the solution color changed from red-orange to green. An infrared spectrum of the green solution revealed the presence of a strong CO stretch at 2079 cm'l. In addition to this stretch, several less intense bands were observed between 2009 cm'1 and 2148 cm'l. An infrared spectrum of the solution taken during the orange stage shows not only the same bands, but also those corresponding to a significant amount of unreacted 19, which implies that the reaction is not complete until the solution has become green. The 1H NMR spectrum of the product (CD3CN) shows three primary resonances at 6 = 3.46 ppm (s), 3.82 ppm (s); m-H, 6.20 ppm (t, 4JP-H = 2.2 Hz). It is worth noting that the IR and NMR spectroscopic characteristics for the green product correspond to those observed for the compound prepared by bulk electrolysis and by chemical oxidation with NO+. Based on the aforementioned infrared and NMR spectral data, one can begin to draw several conclusions as to the identity of this oxidized product. The infrared spectrum of the oxidized product indicates that at least one CO ligand is still present. This is not surprising considering the large number of 219 Ir(III) carbonyl complexes that have been reported.12 The shift of v(CO) to higher energy is consistent with oxidation of the metal center which results in a decrease in the degree of n-back bonding present. Based on the sharp and unshifted natrure of the 1H NMR spectrum, the product is diamagnetic, which is consistent an Ir(III) species. The triplet resonance observed for the meta protons indicates that the phosphines are situated trans to each other. The TMPP ligand resonances are magnetically equivalent and highly symmetric, consistent with either a monodentate bonding mode or a fluxional process that exchanges all potentially equivalent protons. A six coordinate complex, which one would expect if the compound is an Ir(III) species, Mao/fl Me\O/\\ /\ /‘ P P Me- 0 0‘Me Me~ 0 0‘Me 00 Ilr— ’Me 00 Ir—CO OC Me. O \JO-Me V K’O‘ VLO‘ Me’ Me Me cis trans requires that the phosphines are multidentate assuming no additional ligands have been added to the coordination sphere. If this is the case, the molecule must exhibit some fluxional behavior. Indeed, the slightly broadened appearance of the methoxy resonances indicates that an exchange process is operative. The infrared data combined with the NMR data suggest that the observed product is an Ir(III) dicarbonyl species containing either cis or trans carbonyl groups as shown below. Unfortunately, attempts to isolate pure crystalline samples have yielded only oily solids. 220 4. Discussion The primary motivation for investigating the chemistry of TMPP with Rh(I) and Ir(I) olefin complexes was to develop efficient methods for preparing homoleptic TMPP complexes of rhodium and iridium. Others have found that compounds of the type [MI(diene)(solvent)2]1+ are excellent precursors for a number of Rh(I) and Ir(I) phosphine complexes.2 In the present study, TMPP easily displaces the coordinated solvent molecules to form [Rh1(cod)(n2-TMPP)]1+ (1 6). Addition of a second equivalent of phosphine, however, does not result in substitution of the olefin, but instead dealkylates the coordinated TMPP of 16 to produce Rh(cod)(n2-TMPP-O) (17). The strong affinity of the metal for the olefin is further evident by the observation that 17 is inert with respect to substitution with additional TMPP. Reaction of 17 with excess TMPP under refluxing conditions produced only an intractable mixture of products. One possible solution to this problem would be to alter the nature of the coordinated olefin. Replacement of cod with an alkene may encourage dissociation in the absence of the chelate effect. Indeed, reaction of [Rh(CzH5)ZCl]2 with excess TMPP in the presence of KBF4 led to the formation of the same product produced by cobaltocene reduction of [Rh(n3-TMPP)2][BF4]2 (3). Although an analogous ethylene complex is not known for Ir, the corresponding cyclooctene complex [Ir(CgH14)2Cl]2 could be used as a precursor to Ir-TMPP complexes.13 Another approach to synthesizing compounds of the type MI(TMPP)(TMPP- O) (M = Rh, Ir) is by hydrogenation the coordinated diene of 17 and 18 in a donating solvent followed by addition of TMPP. This method has proven effective in the formation of solvated metal phosphine complexes that are active hydrogenation catalysts.14 A third possible route would be to use a non-olefin Rh(I) or Ir(I) precursor such as M(acac)2.15 Presumably, the 221 acetylacetonate ligand would be more susceptible to substitution than cyclooctadiene. Although olefin-phosphine complexes are not novel, compounds 16 - 18 may be useful as synthetic building blocks for other Rh and Ir TMPP complexes. Removal of the diene by either simple substitution or through olefin hydrogenation could lead to the development of a new class of mixed ligand complexes that incorporate the advantages of TMPP with the added reactivity provided by other ligand sets. For example, it may be possible to prepare novel mixed phosphine/thiolate complexes with the general formula, MIH(n2-TMPP)(112-TMPP-0)(SR)2 (M = Rh, Ir), by oxidative addition of alkyl disulfides to 17 - 18 in the presence of TMPP. As an additional point, many Rh and Ir olefin complexes serve as catalyst precursors.14 The combination of a hard oxygen donor with a soft phosphorus donor may induce unusual catalytic properties. As a result, the potential application of these rhodium and iridium systems towards homogenous catalysis is currently under investigation. 16 Previously, we found that [RhI(TMPP)2(CO)][BF4] (7) may be chemically oxidized to yield the paramagnetic Rh(II) compound [Rh(n3- TMPP)2][BF4]2 (3). Oxidation of 7 yield an unstable Rh(II) carbonyl species that immediately undergoes dissociation. By analogy, we expected that oxidation of the iridium dicarbonyl complex 19 might yield an unstable Ir(II) dicarbonyl species that readily dissociates CO to form [IrII(TMPP)2]2+. Clearly, however, CO remains bonded to the metal as evidenced by IR spectroscopy. In any event, it is apparent that we may have reached a limitation of using a containing iridium carbonyl complex as a precursor to homoleptic Ir-TMPP species. Unlike the analogous chemistry of 222 [Rh(TMPP)2(CO)n]1+, CO is more tightly bound to iridium,17 therefore the isolation of [IrH(n3-TMPP)2]2+ may not be feasible using this approach. 10. 223 List of References (a) Bennett, M. A.; Lonstaff, P. A. J. Am. Chem. Soc., 1969,91, 6266. (b) Moers, F. G.; DeJong, J. A. M.; Beaumont, P. M. J. J. Inorg. Nucl. Chem. 1973,35, 1915. (c) Masters, 0.; Shaw, B. L. J. Chem. Soc. (A) 197], 3679. (d) Empsall, H. D.; Heys, P. N.; Shaw, B. L. Transition Met. Chem. 1978, 3, 165. (e) Empsall, H. D.; Hyde, E. M.; Pawson, D.; Shaw, B. L.; J. Chem. Soc., Dalton Trans. 1977, 1292. (f) Mason, R.; Thomas, K. M.; Empasall, H. D.; Fletcher, S. R.; Heys, P. N.; Hyde, E. M. Jones, C. E; Shaw, B. L. J. Chem. Soc., Chem. Comm. 1974, 612. (g) Empsall, H. D.; Hyde, E. M.; Shaw, R. L. J. Chem. Soc., Dalton Trans. 1975, 1690. (h) Empsall, H. D.; Heys, P. N.; McDonald, W. 8.; Norton, M. 0.; Shaw, B. L. J. Chem. Soc., Dalton Trans. 1978, 1119 (a) Schrock, R. R.; Osborn, J. A. J. Am. Chem. Soc. 197], 93, 2397. (b) Shapley, J. R. Schrock, R. R. J. Am. Chem. Soc. 1969,91 , 2816. (c) Green, M.; Kuc, T. A.; Taylor, S. H. J. Chem. Soc. (A) 197], 2334. (d) Haines, L. M.; Singleton, E. J. Chem. Soc., Dalton Trans. 1972, 1891. (e) Haines, L. M. Inor. Nuc. Chem. Lett. 1969, 5, 399. (f) Haines, L. M. Inorg. Chem. 1970, 9, 1517. Herde, J. L.; Lambert, J. C.; Senofl', S. V. Inorg. Synth. 1974, 15, 18. Cramer, R. Inorg. Synth. 1974,15, 14. Bell, F. A.; Ledwith, A.; Sherrington, D. C. J. Chem. Soc. (C) 1969, 2719. (a) Bino, A.; Cotton, F. A.; Fanwick, P. E. Inorg. Chem. 1979, 18, 3558. (b) Cotton, F. A.; Frenz, B. A.; Deganello, G.; Shaver, A. J. Organomet. Chem. 1973, 50, 227. TEXSAN - TEXRAY Structure Analysis Package, Molecular Structure Corporation, 1985. DIRDIF: Direct Methods for Difference Structure, An Automatic Procedure for Phase Extension; Refinement of Difference Structure Factors. Beurskens, R. T. Technical Report 1984. (a) Dunbar, K. R.; Haefner, S. C.; Burzynski, D. J. Organometallics 1990, 9, 1347 . (b) Dunbar, K. R.; Haefner, S. C.; Bender, C. J. Am. Chem. Soc., 1991, 113, 9540. (c) Dunbar, K. R.; Haefner, S. C.; Swepston, P. N. J. Chem. Soc., Chem. Commun. 199], 460. (a) Garrou, P. E. Inorg. Chem. 1975, 14, 1435. (b) Garrou, P. E.Chem. Rev. 198], 81, 229. 11. 12. 13. 14. 15. 16. 17. 224 (a) Johnson, C. E.; Eisenberg, R.; Evans, T. R.; Burberry, M. S. J. Am. Chem. Soc. 1983, 105, 1795. (b) Fordyce, W. A.; Crosby, G. A. Inorg. Chem. 1982, 21 , 1455. (c) Fordyce, W. A.; Crosby, G. A. Inorg. Chem. 1982,21, 1023. (d) Fordyce, W. A.; Rau, H.; Stone, M. L.; Crosby, G. A. Chem. Phys. Lett. 198], 77, 405. (e) Andrews, L. J. Inorg. Chem. 1978, 17, 3180. (f) Geoffroy, G. L.; Isci, H.; Litrenti, J. Mason, W. R. Inorg. Chem. 1977, 16, 1950. (g) Brady, R.; Flynn, B. R.; Geoffroy, G. L.; Gray, H. B.; Peone, J. Jr.; Vaska, L. Inorg. Chem. 1976, 15, 1485. (h) Geoffrey, G. L.; Wrighton, M. 8.; Hammond, G. 8.; Gray, H. B. J. Am. Chem. Soc. 1974, 96, 3105. (i) Brady, R.; Miller, W. V.; Vaska, L. J. Chem. Soc., Chem. Commun. 1974, 393. Leigh, G. J.; Richards, R. L. in Comprehensive Organometallic Chemistry; Wilkinson, G.; Stone, F. G. A.; Abel, E. W., eds.; Pergamon: New York, 1982; vol. 5, p 553. (a) Herde, J. L.; Lambert, J. C.; Senoff, S. V. Inorg. Synth. 1974, 15, 18. (b) Herde, J. L.; Senoff, Inor. Nuc. Chem. Lett. 197], 7, 1029. (a) Crabtree, R. Acc. Cham. Res. 1979, 12, 331. (b) Schrock, R. R.; Osborn, J. A. J. Am. Chem. Soc. 1976, 98, 2134. (c) Schrock, R. R.; Osborn, J. A. J. Am. Chem. Soc. 1976,98, 2143. Add halpern and catalyst precursor Fackler, J. P. Prog. Inorg. Chem. 1966, 7, 361. Horvath, I. work in progress. Ibers, J. A.; Lilga, M. A. Inorg. Chem. 1984,23, 3538. CHAPTER IX CONCLUSION 225 226 The results presented in this dissertation establish the use of tris(2,4,6-trimethoxyphenyl)phosphine as an excellent supporting ligand for the stabilization of divalent rhodium species. The mononuclear Rh(II) complex, [Rh(n3-TMPP)2][BF4]2 (3) was isolated in high yield from a reaction of the solvated dinuclear Rh(II) species [Rh2(MeCN)10][BF4]4 with TMPP. [Rh(n3-TMPP)2][BF4]2 (3 ) represents the first crystallographically characterized mononuclear Rh(II) complex. The remarkable stability of this metalloradical species provided us with an opportunity to study the chemistry of mononuclear rhodium in an unusual oxidation state. The combination of both hard and soft donor groups engenders the phosphine with the ability to stabilize a metal atom in a variety of coordination geometries and oxidation states. Although considered sterically encumbering, the ligand is nevertheless quite flexible as evidenced by the variety of fluxional and isomerization processes in which have been observed for metal complexes of TMPP. This flexibility has rendered it possible to access different oxidation states of the rhodium center while maintaining the same ligand set. Each of the complexes described in this study, exhibit one or more chemically accessible redox processes. The chemical relationship between the rhodium-TMPP complexes in this study is illustrated in Figure 37. The degree of methoxy interaction varies according to the rhodium oxidation states. Rather than serving to enforce a particular geometry on the metal center, the ligand adjusts to the available coordination sites and the electronic requirements of the metal center (Figure 38). This ability is readily apparent when one examines the structural changes in going from Rh(I) to Rh(III) species. As the metal center is oxidized, the hapticity of the phosphine ligand increases from n1 to n3 in response to the increased need for electron density at the metal center. «>' .1“; 227 4.2002 .2225“: 0.325.388 5222.255 .~.oN:o.s=m20£>5 2.5235 s c.0035 doc .358 8.3236335 25.938 «223223 .o.£-.~.om=o .mzésg538322053 sauna60985388095 .«_ommo.oo«63 .maumoésao 82522.": 23811 .680 £3 .s...zom.2aE .003-222.333.2805: 292.2293 .5255: s .222. 9c 22.255.231.825 .93.. .3035 .3852”: s 981222.43on: .05 .2002 638 3 22s Ea 3: :c a: :Egesissgfisé :Eamzoxagifi .s:o-&2e-m:xmg-ms=_5:8 cw: +~HNAOIAHAWEIMP5~=§H|UQ~§ E a: a: 9: a: so a: As Inasmzoxooxamgifi amazmzovNEmEV—EE :zodmzexmmgxifisoa $98926.-."cmedE—imwud its #3 ~25 8V 2.25 A5 A8 A3 IEOOEAQEVEE lllmvl zzoovuflmzeifi lit Iv JEmEdEiE :5 ixmgdsfisfi >N . : ; 2.: 2838225525 E: a: +HHAgUXHHAHEEnNCVMF—mfl j :+~HNCH~HVAMHOUV—£MHH: 2:5 £28326 Lo§2ozv=sug «.8335 we 8.22988 mg 8 223:8 oamfigm Sm 0.33% 228 5-22.: Es “E22. .5. 888 32:8. as - H: .8 8.52 n: «c. «SO x O“ZIAV QMO\QE m\\~‘ 0!; 21w; W2 .25 o O m.2 / o: O \C 0519 z a: O u z 0 82253 04.5.): \O 02 c c c 02, N _ O ~— a: 025 a...) @1522 ..TJ2 a: \ S‘O o \sémlz Ow: ’02 ¢\ \O 573 a: 22:55 has: gas... no mono—2 mEucom m 229 A critical point of the chemistry of TMPP was underscored in this research, viz, that although [Rh(n3-TMPP)2][BF4]2 (3) is coordinatively saturated, the hemilabile nature of the ether groups allows for facile and often reversible substitution chemistry. It was found that the incoming ligand must possess certain electronic requirements as the Rh(II) phosphine complexes are resistant to attack by moisture or oxygen but are quite susceptible to n-acids such as CO, CNR and NO. The ability of the ether groups to participate in such "arm on, arm off ' type of mechanism is critical - for the future development of these complexes as viable systems for catalytic or stoichometric transformations of small substrates. The work described in chapter 5 demonstrates the possibility of exploiting the labile nature of the pendent methoxy substituents for the development of molecule-based chemical sensors. The future design of other selective chemical sensors based on complexes of TMPP appears to be very promising. The complex chemistry of [Rh(n3-TMPP)2][BF4]2 (3) with carbon monoxide provided valuable insight into the subtle electronic influences that govern the relative stabilities of the +1, +2 and +3 oxidation states. The reversible interaction of carbon monoxide with [Rh(n3-TMPP)2][BF4]2 (3) was found to proceed through a series of redox reactions brought about by the initial disproportionation reaction between 3 and an intermediate Rh(II) carbonyl species. By reducing the n-acceptor strength of the incoming ligand through the use of alkyl isocyanide ligands, the redox pathway was shutdown and stable isocyanide adducts of 3 were isolated. At the time of their discovery, the complexes [RhH(TMPP)2(CNR)2][BF4]2 (R = But, Pri) represented the first mononuclear organometallic Rh(II) species sufficiently stable enough to be spectroscopically and crystallographically characterized. 230 The stability of these complexes provides for a rare opportunity to investigate the chemistry of an organometallic radical system. For the purpose of future directions, it should be pointed out that a potential drawback to the use of TMPP as a supporting ligand is its documented susceptibility to nucleophilic attack at a coordinated methoxy substituent. This results in dealkylation of the coordinated ether group and formation of a metal-phenorn'de bond. The tendency of these systems to dealkylate is highly dependent on the electrophilicity of the metal center. This point was illustrated by the relative stabilities of the molecular cations [RhH(n3-TMPP)2]2+ (3) and [RhIH(n3-TMPP)2]3+ (4) towards dealkylation. As the electrophilicty of the metal center was increased, the ease of dealkylation also increased. It is expected that such behavior will be particularly troublesome for the more electrophilic early transition metals; as a result, this will necessarily limit the use of the free phophine form of TMPP to lower valent early and late transition metals. One of our principlal goals prior to undertaking this chemistry was to the develop the coordination and organometallic chemistry of odd-electron systems. The work presented here has demonstrated that TMPP provides the proper combination of kinetic and thermodynamic stability for the isolation of mononuclear Rh(II) complexes. Clearly, based on the demonstrated ability of TMPP to stabilize Rh(II) metalloradicals, extension of this work to other d7 metal systems is plausible. Indeed, recent work in our laboratories has led to the successful isolation of the Ni(III) complex, [Ni(TMPP-0)2]1+.1 Unfortunately, initial efforts to synthesize analogous Ir(II) species have proven unsuccessful. This is attributed, in part, to the lack of a suitable precursor such as a solvated [Ir2]4+ species. Undoubtedly, further work in this area will ultimately result in the isolation of stable Ir(II)-TMPP systems. 231 As for complexes of Pd, Pt and Ru, nitrile complexes of these metals are known213’4 and should provide excellent precursors for the synthesis of d8 and d6 metal TMPP complexes. These intermediates, in turn, may be either oxidized or reduced to the corresponding paramagnetic d7 species. In related work, Gladfelter et al. have shown that paramagnetic Ru(I) carbonyl complexes are accessed by chemical ozddation of zero-valent Ru(PR3)2(CO)3 species (R: phenyl, benzyl, p-tolyl, cyclohexyl).5 Similar d7 carbonyl species have also been isolated and crystallographically characterized for Re(0).6 By analogy, the preparation of similar metalloradicals of the general formula [M(TMPP)2(CO)x]n+ is feasible. These odd-electron carbonyl complexes will be especially intriguing in light of the rich chemistry already documented for TMPP complexes of Rh(I), Ir(I) and Mo(0) carbonyl compounds. 232 List of References 1. Dunbar, K. R.; Quillevéré, A submitted for publication. 2. Thomas, R. R.; Sen, A. Inorg. Synth. 1989, 26, 128. 3. De Renzi, A.; Panunzi, A.; Vitagliano, A.; Paiaro, G. J. Chem. Soc., Chem. Commun. 1976, 47 . 4. (a) Rapaport, I.; Helm, L.; Merbach, A. E.; Bernhard, P.; Ludi, A. I norg. Chem. 1988, 27, 873. (b) Bown, M.; Fontaine, X. L. R.; Greenwood, N. N.; Kennedy, J. D.; Thornton-Pett, M. J. Chem. Soc., Dalton Trans. 1987, 1169. (c) Kolle, U.; Flunkert, G.; Gérissen, R.; Schmidt, M. U.; Englert, U. Angew. Chem. Int. Ed. Engl. 1992, 31 , 440. 5. Sherlock, S. J .; Boyd, D. C.; Moasser, Gladfelter, W. L. Inorg. Chem. 199], 30, 3626. 6. (a) Walker, H. W.; Rattinger, G. B.; Belford, R. L.; Brown, T. L. Organometallics 1983,2, 775. (b) Crocker, L. S.; Heinekey, D. M.; Schulte, G. K. J. Am. Chem. Soc. 1989,111, 405. APPENDICES APPENDIX A PHYSICAL MEASUREMENTS APPENDIX A PHYSICAL MEASUREMENTS 1. Infrared Spectroscopy Infrared spectra were recorded on a Perkin- Elmer 599 or a Nicolet 740 FT-IR spectrophotometer. Solid state spectra were measured as Nujol mulls between CsI plates. Solution spectra were recorded using CaF2 solution cells. 2. NMR Spectroscopy 1H NMR spectra were measured either on a WM 250-MHz Bruker spectrometer with an ASPECT 3000 computer, a Gemini 300-MHz, Varian 300-MHz or Varian 500-MHz spectrometer. Chemical shifts were referenced relative to the residual proton impurity of the deuterated solvent: d2-methylene chloride, 5.32 ppm; d3-acetonitrile, 1.93 ppm; de-acetone, 2.04 ppm; dl-chloroform, 7.24. 13'C{1H} NMR spectra were recorded on a Varian BOO-MHz spectrometer operating at 73.1 MHz and were referenced relative to the 13C solvent resonance. 31P{1H} NMR spectra were obtained on a Varian 300-MHz spectrometer operating at 121.4 MHz. Chemical shifts were referenced relative to an external standard of 85% H3PO4. Positive chemical shifts were reported downfield relative to H3PO4. 3. EPR Spectroscopy X-band EPR spectra were obtained using a Bruker ER200D spectrometer. To obtain an accurate measure of g values and line widths, a Bruker ER035M NMR Gaussmeter and a Hewlett-Packard 5245L frequency counter (with a 3-12 GHz adapter) were used to measure magnetic field strength and microwave frequency, respectively. 4. Electronic Absorption Spectroscopy Electronic absorption spectra were measured on a Hitachi U-2000 or a Cary 17 spectrophotometer. 234 5. Electrochemistry Electrochemical measurements were performed by using an EG&G Princeton Applied Research Model 362 scanning potentiostat in conjunction with a BAS Model RXY recorder. Cyclic voltammetry experiments were carried out at 22+2°C using 0.1 M tetra-n-butylammonium tetrafluoroborate (TBABF4) as a supporting electrolyte, unless noted otherwise. Measurements were made at a scan rate of 200 mV/s using a platinum disk working electrode and a Ag/AgCl reference electrode. E1/2 values, determined as (Ep,a+Ep’c)/2, were referenced to a Ag/AgCl electrode and uncorrected for junction potentials. Under the same experimental conditions, the szFe / CpZFe+ couple occurred at Em = +0.50 V. 6. Magnetic Susceptibility Variable temperature and field magnetic susceptibility measurements were carried out on a 10 KG BTI Superconducting Quantum Interference Device (SQUID) at Michigan State University. Additional solid state magnetic susceptibility measurements were determined at room temperature by using a J ohnson-Matthey magnetic susceptibility balance. Solution magnetic susceptibility measurements were carried out by application of the Evans method on a Bruker 250-MHz or Varian 300-MHz spectrometer. 7. Mass Spectrometry Fast Atom Bombardment (FAB) mass spectrometry studies were performed on a JEOL HX double-focusing mass spectrometer housed at the National Institutes of Health / Michigan State University Mass Spectrometry Facility. 8. Elemental Analysis Elemental analyses were performed at Galbraith Laboratories, Inc. APPENDIX B CHEMISTRY OF TRIS(2,4,6-TRIIVIETHOXYPHENYL)PHOSPHINE WITH GROUP VI METALS APPENDIX B CHEMISTRY OF TRIS(2,4,6-TRIMETHOXYPHENYL)PHOSPHINE WITH GROUP VI METALS 1. Introduction As our eventual goal in this research is to establish that TMPP, with its labile metal-ether interactions, is an excellent ligand for the design of reactive metal centers, we set out to prepare a complex that would allow us to probe the small molecule binding properties of a TMPP supported metal. Based on our newly acquired knowledge of the bonding capabilities of the ligand, we predicted that it would be possible to prepare compounds of general formula (TMPP)ML3 for metals that ordinarily exhibit octahedral structures. A convenient backdrop for our work is the elegant research of Kubas and co-workers who demonstrated that highly donating, bulky phosphine ligands stabilize the five-coordinate complexes M(CO)3(PR3)2 M(=Mo, W; R=CGH11, i-C3H7) which readily add molecular H2 and N 2.1 Since no examples of four-coordinate complexes of general formula M(CO)3(PR3) have been reported, we set out to prepare such a monophosphine derivative with TMPP with the expectation that the product would exhibit unusual properties and reactivity. Herein, we report the synthesis of the novel fluxional molecule (n3-TMPP)M0(CO)3 (TMPP = tris(2,4,6- trimethoxyphenyl)phosphine) from a reaction between (116-C 7H8 )Mo(CO)3 and TMPP. In addition, the chemistry of TMPP with partially solvated mononuclear and dinuclear molybdenum halide complexes is also presented. 235 -_— 236 2. Experimental A. Synthesis All reactions were carried out under an argon atmosphere by the use of standard Schlenk-line techniques unless otherwise stated. The phosphine, tris(2,4,6-trimethoxyphenyl)phosphine (1), was prepared as described in Chapter II. The starting material, (n6-C7H8)M0(CO)3, was purchased from Strem and used as received. M02C14(MeCN)4 and MoCl3(THF)3 were prepared using standard literature procedures.2’3 The tungsten starting material, (nG-C7H8)W(CO)3, was prepared following the literature procedure described by Kubas.4 (1) Preparation of (n3-TNIPP)M0(CO)3 (20) In a typical synthesis, a quantity of (n6-C7H8)Mo(CO)3 (0.300 g, 1.10 mmol) and TMPP (0.587 g, 1.10 mmol) was dissolved in 10 mL of benzene that had been thoroughly purged with argon in order to remove dissolved N2(g). The resulting suspension was stirred at r. t. for ca. 12 h. The orange supernatant was decanted and the yellow precipitate was washed with benzene (4 x 5 mL). The yellow product was recrystallized by dissolution in CH2012 (10 mL), followed by slow addition of diethyl ether (10 mL). The solution was decanted and the yellow crystalline solid was washed with diethyl ether (3 x 5 mL) and dried in vacuo; yield, 0.624 g (80 %) IR (Nujol, CsI) cm'l: v(CO), 1914.5 vs. 1775 vs. 1791 vs. IR (CHZCIZ) cm‘l: v(CO), 1921 vs, 1799 vs, 1782 vs. IR (CGH6) cm'1: v(CO), 1930 vs. 1798 s. 1H NMR (CD2C12, 20°C) 8 ppm: ~OCH3, 3.55 (br), 3.81 (s); m-H, 6.15 (br). 1H NMR (06D5CD3, -60°C) 6 ppm: o-OCH3, 2.80 (s, 6H), 3.19 (s, 6H), 3.81 (s, 6H); p- OCH3, 3.33 (s, 6H), 3.44 (s, 3H); m-H, 5.77 (s, 2H), 5.86 (s, 2H), 6.01 (br, 2H). 1H NMR (CDZCI2, -4000) 5 ppm: o-OCH3, 3.49 (s, 12H), 4.36 (s, 6H); p-OCH3, 237 3.77 (s, 6H), 3.82 (s, 3H); m-H, 6.02 (s, 2H), 6.07 (d, 4Jp_H= 3.9 Hz, 2H), 6.15 (dd, 2H). 31P NMR(CD2012, 20°C) 8 ppm: -1.9 (s). (2) Reaction of TMPP with (n 6-C7H3)W(CO)3 To a mixture of (n6-C7H8)W(CO)3 (0.215 g, 0.597 mmol) and TMPP (0.318 g, 0.597 mmol) was added 5 mL of diethyl ether. The resulting suspension was stirred for 12 h at r.t. During this time, a yellow solid precipitated from the solution. The yellow product was collected by filtration, washed with benzene (2 x 5 mL) and diethyl ether (4 x 10 mL), and dried under vacuum; yield of crude precipitate 0.333 g. IR (Nujol, CsI) cm'l: v(CO), 1910 vs. 1769 vs. IR (CHZClg) cm'1:v(CO), 1912 vs, 1789 vs, 1775 vs. (3) Reaction of TMPP with MoCl3(THF)3 ‘ A solution of TMPP (0.382 g, 0.717 mmol) in 10 mL of THF was added dropwise to a solution of M0C13(THF)3 (0.300 g, 0.717 mmol) in 5 mL of THF at 0°C. The reaction was stirred for ca. 12 h, during this time the solution was allowed to slowly warm to r. t. The resulting red solution was filtered through a Celite plug and then evaporated to yield a red solid. The solid was washed with diethyl ether (3 x 5 mL) and dried under reduced pressure; yield of red solid, 0.348 g. 1H NMR (CD3CN) 8 ppm: -OCH3, 3.51 (s), 3.82 (s); m-H, 6.14 (d, 4Jp_H = 5.7 Hz). (4) Reaction of TMPP with M02014(MeCN)4 (i) 4 equivalents. In a typical reaction, a mixture of M02C14(MeCN)4 (0.250 g, 0.502 mmol) and TMPP (1.07 g, 2.01 mmol) was dissolved in 25 mL of MeCN at r.t. After 3 h, the solution had become green in color. The solvent was evaporated under reduced pressure to yield a green solid that was washed with 20 mL of benzene. The colorless benzene washing was evaporated to give a white powder (0.41 g) that was shown to be TMPP by 1H NMR spectroscopy. The green residue was further washed with 10 mL of 238 diethyl ether and dried under vacuum; yield of crude solid, 0.48 g. A 1H NMR spectrum of the crude product reveals the presence of only [TMPP-HT: (ii) 6 equivalents. In a typical reaction, a solution of M02C14(MeCN )4 (0.200 g, 0.402 mmol) and TMPP (1.28 g, 2.40 mmol) in 20 mL of MeCN was refluxed for 4 days. During this time, the solution color changed from green to dark brown with no further change throughout the reaction. The solution was then filtered through a Celite plug and evaporated to a residue. The brown solid was washed with benzene (2 x 10 mL) and diethyl ether (2 x 10 mL) and dried under vacuum; yield of brown solid, 0.766 g. 1H NMR spectroscopy showed that the solid was primarily comprised of [TMPP-CH3]+. B. X-ray Crystallography The structure of (n3-TMPP)Mo(CO)3 (20) was determined by application of general procedures that have been fully described elsewhere.5 Geometric and intensity data were collected on a Nicolet P3/F diffractometer with graphite monochromated MoKoc Ova: 0.7107 3 A) and were corrected for Lorentz and polarization effects. All calculations were performed on a VAXSTATION 2000 computer. Data reduction and refinement were performed using the programs from the Enraf-Nonius Structure Determination Package (SDP). Crystal parameters and basic information pertaining to data collection and structure refinement are summarized in Table 17 (l) (113-TMPP)M0(CO)3 (20) 0 CH2012 (i) Data Collection and Reduction. Crystals of (n3-TMPP)M0(CO)3 (20) - CH2C12 were obtained from slow diffusion of diethyl ether into a CH2012 solution of 20. A single crystal with approximate dimensions 0.60 x 0.40 x 0.15 mm3 was selected and mounted onto the tip of a glass fiber with epoxy cement. Least-squares refinement of 25 carefully centered reflections in the 239 Table 17. Summary of crystallographic data for (n3-TMPP)Mo(CO)3 . CH2C12 (20). Formula Formula weight Space group a, A b, A c, A or, deg [3, deg 7, deg V, A3 Z dcalc, g/cm3 u (Mo Ka), cm'1 Temperature, °C Ra wa Quality-of-fit indicatorc MoPC12012031H35 801.92 P2 Va 17. 155(8) 1.2019(4) 16.985(6) 90 9569(3) 90 3485(2) 4 1.493 6.219 22:2 °C 0.069 0.084 2.09 aR=Z I lFol' chll/leol bRw = [M I F0 I— I Fc I)2/2w tro I211/2; w = 110% [PO I) cQuality-of-fit = [2w( IFO I - IFc[)2/(Nobs-Nparameters)]1’2 240 range 15 < 20 < 25° revealed that the compound (n3-TMPP)M0(CO)3 - CH2C12 crystallized in a monoclinic space group P21/a with a = 17 .155(8) A, b = 12.019(4) A, c = 16.985(6) A, [3 = 9569(3) A, V = 3485(2) A3. Axial photographs confirmed the monoclinic symmetry of the cell. Intensity measurement of three representative reflections at regular intervals throughout data collection revealed that a 41.7 % loss in diffraction intensity had occurred. A linear decay correction was applied to the data using the program CHORT in SDP. A Nicolet. P3/F diffractometer was used to collect ‘ 3845 unique data in the range 4 S 20 S 43° at 22 i 2°C; 2243 data with F02 > 36(F02) were used in the refinement. (ii) Structure Solution and Refinement. The position of the molybdenum atom was successfully located by application of the direct methods program in SHELXS-86. The remaining non-hydrogen atoms were located through successive least-squares refinements and difference Fourier maps. After the refinement had successfully converged with isotropic thermal parameters, an absorption correction based on the program DIFABS was applied to the data.6 In the end, after anisotropic refinement of 424 parameters, residuals of R = 0.069 and RW = 0.084 were obtained. The quality-of-fit index was 2.09 and the largest shift/esd was 0.08. A final difference Fourier map revealed that no peak remained over 0.86 e'/A3 in height. 3. Results A. Synthesis and Characterization of (113-TMPP)M0(CO)3 (20) One equivalent of TMPP reacts smoothly with (n3-C7H8)Mo(CO)3 at room temperature in benzene over the period of 6 hours to produce a yellow microcrystalline compound formulated as (n3-TMPP)M0(CO)3 (20) in 80% 241 yield. The product is insoluble in most common solvents except toluene and benzene in which it is sparingly soluble and dichloromethane in which it is very soluble but prone to decomposition over long periods of time. Yellow solutions of 20 are very air sensitive, turning brown immediately upon exposure to air. The solid and crystalline forms decompose slowly over the period of several hours. The solution properties of (n 3-TMPP)Mo(CO)3 attest to its high reactivity as it easily converts to Mo(CO)3(NCCH3)3 in acetonitrile and is extremely air sensitive. The 1H NMR spectrum of 20 revealed that an intramolecular exchange process involving the ortho-methoxy groups is occurring at room temperature. Variable temperature 1H NMR data were obtained in d8-toluene and CD2C12 over the range 420°C and -60°C and the results clearly indicate that all three rings are participating in a low energy fluxional process (Figure 39). The low temperature limiting spectrum at -60°C in d8-toluene exhibits eight distinct resonances which integrate in accordance with the magnetically inequivalent meta, ortho and para groups observed in the solid state structure. At temperatures above -60°C the spectral features broaden and gradually collapse in a non-symmetrical manner due to a dynamic exchange of interacting and non-interacting ortho- methoxy groups. Concomitantly the para and meta regions broaden and eventually coalesce at ca -15°C. Similar behavior is observed in CD2012 , although in this solvent the low temperature limiting spectrum shows only one broad resonance for the non-interacting ortho-methoxy groups at 5 = + 3.49 ppm (Figure 40); Attempts to obtain spectra at higher temperatures were thwarted by the thermal instability of the complex. 242 5833-3. 5 88 20952352555 558% mzz E 8558383 535.5> .3 9:6:— End Lll . own mun . cue . .mnv. . cum. mum . cum mum 8- 1<