arm? . I \, ‘s _ 13$qu uému. RIC . 1k ’3 J .15.: ‘_ L‘ fix}? 4. fififié‘tk 3?. S 3‘ .r .mwfi «5... .3- t n...“ u}! 2.0.3:}. ... b. . i. _ I l\..i1\ c!!! 5.1.1 l}; 1513.. a I)». 1030 LIBRARY MiC“.‘;¢' State University This is to certify that the dissertation entitled DEVELOPMENT OF TITANIUM PYRROLYL HYDROAMINATION CATALYSTS AND URANIUM PYRROLYL COMPLEXES presented by Douglas L. Swartz II has been accepted towards fulfillment of the requirements for the Ph.D degree in Chemistry ___,, , , . 2 Major Professor's Signature MSU is an Affirmative Action/Equal Opportunity Employer u-..--. PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 K:lProleoc&Pres/ClRC/Dateom.indd DEVELOPMENT OF TITANIUM PYRROLYL HYDROAMINATION CATALYSTS AND URANIUM PYRROLYL COMPLEXES By Douglas L. Swartz II A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree DOCTOR OF PHILOSOPHY Chemistry 2009 ABSTRACT DEVELOPMENT OF TITANIUM PYRROLYL HYDROAMINATION CATALYSTS URANIUM PYRfiOqfYL COMPLEXES By Douglas L. Swartz II The primary focus of this thesis is the design and development of pyrrolyl based titanium hydroamination catalysts with applications towards multi-component coupling reactions involving a primary amine, alkyne, and isonitrile. Hydroamination is an atom efficient process for the production amines and imines from the formal addition of an N- H bond across a C—C unsaturated bond. Titanium-catalyzed hydroamination has seen an explosion of activity and has led to new methodologies in a variety of C—N containing molecules. The first chapter briefly discusses the types of ancillary ligands employed for titanium hydroarnination catalysts with the main focus being on the development of titanium pyrrolyl complexes for the development of C—N bond forming reactions. Since the first successful hydroamination of a primary amine and alkyne by a titanimn pyrrolyl complex our goal has been to optimize the most promising catalysts that carry out these reactions to expand the scope of this methodology. Chapters 2 — 4 discuss the types of electronic features and steric profiles in catalyst design that encourage these useful C-N bond forming reactions. Ligand isomerization in titanium dipyrrolylmethane complexes is common due to the different bonding hapticities the ligand can adopt. Methods for altering this barrier may provide clues to the active species in catalysis and allow control of complex structures. Chapter 5 discusses the effects 5,5-substitution has on dipyrrolylmethane ligand isomerization and the parameters for pyrrolyl exchange. Since the discovery of uranium bis(imido) analogues of the uranyl ion, U022”: actinide chemists have been intrigued with the bonding and reactivity of this functional group. Pyrrolyl ligands have proven to be a useful class of ancillary ligands for transition metals, however their employment in actinide chemistry is relatively scarce. The synthesis, structure, and reactivity of uranium bis(imido) dipyrrolylmethane complexes are discussed in Chapter 6. To my wife, I hope it was all worth it ACKNOWLEDGEMENTS First and foremost I have to thank wife, Melesa, for all the love and support she has given to me over the past 11 years we have been together, especially the past 5 years during my time as a graduate student at Michigan State. I also want to thank my son, Lucelus, for being the best part of my day everyday when I get home, greeting me with the best hug and smile a dad can ever get. I would like to thank my advisor, Aaron Odom, for all his patience, guidance and support during my time at Michigan State. Working under his direction provided me with wonderful opportunities that have lead to a satisfying graduate career. Among the many opportunities I was able to take part in, working with Dr. Jim Boncella at Los Alamos National Lab was an absolute pleasure and privilege. I would like to thank Jim for all his advice, patience, support he gave to my family and I while we were in Los Alamos. I would also like to acknowledge all my past and present group members: Jim, Bala, Kapil, Sanjukta, Sam, Supriyo, Steve, and Eyal. I would like to give Supriyo a special “thank you” for being such a wonderful friend and colleague. I will miss working alongside him, “Absolutely.” I would like to acknowledge Professor Mitch Smith, Professor Jim McCusker, and Professor Bill Wulff for serving on my committee, for their guidance, and teaching me a degree of professionalism in presenting my research. A special thanks goes to Dan Holmes for assisting me with all the NMR experiments I had to conduct throughout my graduate career. Lastly, I would like to thank my parents for all the love, support and faith they put in me that helped me succeed in so aspects of life. I hope that I have made them proud. Douglas L. Swartz II vi TABLE OF CONTENTS List of Tables .................................................................................................................... ix List of Figures ..................................................................................................................... x List of Abbreviations ....................................................................................................... xiii List of Schemes ................................................................................................................. xv 1. Titanium pyrrolyl catalysts in C—N bond formation ................................................. l 1.1 Introduction .............................................................................................................. 1 1.2 Hydroamination with titanium Cp complexes .......................................................... 3 1.3 Hydroamination with titanium amido complexes .................................................... 5 1.4 Hydroamination with titanium amidate complexes .................................................. 7 1.5 Hydroamination with titanium pyrrolyl complexes ................................................. 8 1.6 Iminoamination ...................................................................................................... 16 1.7 Conclusions ............................................................................................................ 18 1.8 References .............................................................................................................. 19 2. Synthesis, structure, and hydroamination kinetics of 2,2’- diaryldipyrrolylmethane and bis(2-arylpyrrolyl)titanium complexes .................. 21 2.1 Introduction ............................................................................................................ 21 2.2 Titanium pyrrolyl hydroamination ......................................................................... 22 2.3 Hyrdoamination kinetics with titanium dipyrrolylmethane derivatives ................. 26 2.4 Conclusion .............................................................................................................. 41 2.5 Experimental .......................................................................................................... 43 2.6 References .............................................................................................................. 57 3. Synthesis, structure, and hydroamination reactivity of electron deficient titanium complexes hearing bis- and mono(pyrrolyl) ligands ............................................... 59 3.1 Introduction ............................................................................................................ 59 3.2 Results and discussions .......................................................................................... 60 3.3 Conclusion .............................................................................................................. 70 3.4 Experimental .......................................................................................................... 71 3.5 References .............................................................................................................. 81 4. Synthesis, structure, and hydroamination kinetics of 3,3’- diaryldipyrrolylmethane titanium complexes .......................................................... 82 4.1 Introduction ............................................................................................................ 82 4.2 Results and discussions .......................................................................................... 83 4.3 Conclusion .............................................................................................................. 90 4.4 Experimental .......................................................................................................... 91 vii 4.4 Experimental .......................................................................................................... 91 4.5 References ............................................................................................................. 104 5. Effects of 5,5-substitution on dipyrrolylmethane ligand isomerization ................ 105 5.1 . Introduction .......................................................................................................... 105 5.2 Ligand isomerization study ................................................................................... 109 5.3 Conclusion ............................................................................................................. 115 5.4 Experimental ......................................................................................................... 116 5.5 References ............................................................................................................. 122 6. Uranium (VI) bis(imido) pyrrolyl complexes: synthesis, structure, and reactivity ............................................................................................................. 123 6.1 Introduction ........................................................................................................... 123 6.2 Uranium bis(imido) dipyrrolylmethane complexes .............................................. 126 6.3 Conclusion ............................................................................................................. 135 6.4 Experimental ......................................................................................................... 136 6.5 References ............................................................................................................. 141 7. Appendix A Kinetic plots for selected catalysts from Chapter 2 ............................................... 143 B Kinetic plots for selected catalysts from Chapter 3 ............................................... 149 C Kinetics plots for selected catalysts from Chapter 4 ............................................. 152 D NMR spectra for selected compounds ................................................................... 155 viii Table 1.1 Table 1.2 Table 1.3 Table 1.4 Table 1.5 Table 2.1 Table 2.2 Table 3.1 Table 4.1 Table 4.2 Table 5.1 LIST OF TABLES Representative hydroamination results with 10 mol% Ti(NMe2)2(dpma) (8) ............................................................................... 10 Representative hydroamination results with Ti(NMe2)2(dpm) (9) as the catalyst ................................................................................................ 12 Rate constants for Ti(NMe2)2(dpma) (8), Ti(NMe2)2(dpm) (9), TiCNMe2)2(pyrr-Cp) (10) for the hydroamination of aniline and 1- phenylpropyne ........................................................................................... 1 3 Respresentative hydroamination results with Ti(NMe2)2(dap)2 (11) at 10 mol% catalyst loading .......................................................................... 15 Results of 3-component coupling with Ti(NMe2)2(dpma) (8) at 10 mol% catalyst loading at 100 °C in toluene .............................................. 17 Observed rate constants for catalysts TiCNMe2)2(dpm) (9), Ti(NM62)2(dpm;n:l%2). Ti2. . , . mes T1(NMe2)2(pyrr )2 (14), T1(NMe2)2(pyrr )2 (15) ...................... 35 Comparison of rate constants for hydroamination of the bis(pyrroly1) catalysts ..................................................................................................... 37 Comparison of Ti(NMe2)2(pyr2_CF3—4C6F5) (20), Ti(NMe2)2(pyrrmes)2 (l3), and Ti(indenyl)2Me2 (21) ................................................................. 69 Comparison of hydroamination catalysis rate constants . 3,5—CF3 . 4—CF3 T10‘1M62)2(Pyrr )2 (14) and T1(N1\/It‘=2)2(pyrr )2 (16) from Chapter 2 ................................................................................................... 84 Representative catalysis rates for Ti(NMe2)2(dpm) (9), Ti(NMe2)2(NHMe2)(3-dpm3’5—CF3) (22) and Ti(NMe2)2(NHMe2)(3- dme6F3) (23) ........................................................................................... 89 Parameters for pyrrolyl exchange for 9, 24, and 25 at 25 °C .................. 113 ix Figure 1.1 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 Figure 3.1 Figure 3.2 Figure 4.1 Figure 4.2 Figure 5.1 Figure 5.2 Figure 5.3 Figure 5.4 LIST OF FIGURES Titanium amidate complex ......................................................................... 7 ORTEP structure from single-crystal X-ray diffraction of Ti(NMe2)2(dpm3’5—CF3) (13) ................................................................... 29 ORTEP structure from single-crystal X-ray diffraction on Ti(NMe2)2(pyrrmes)2(15) ......................................................................... 31 ORTEP structure from single-crystal X-ray diffraction on Ti(NMe2)2(pyrr3’5‘CF3)2 (14) .................................................................. 32 Plot of catalyst concentration of 9 versus pseudo first-order rate constant for kinetic reaction conditions .................................................... 33 Representative plot of ln[l-phenylpropyne] vs time with complex 9 as the hydroamination catalyst .................................................................. 34 ORTEP structure from singgle5 03;? X-ray diffraction on Ti(NMe2)2(NHMe2)(3-pyr ’ )2 (19) with thermal ellipsoids at 50% probability level ........................................................................................ 62 ORTEP structure from single crystal X-ray diffraction on Ti(NMe2)2(NHMe2)(3-pyr3’S'CF3)2 (20) with thermal ellipsoids at 50% probability level ........................................................................................ 65 Solid-state structure of Ti(NMe2)2(dpm) (9) ............................................ 82 Solid-state structure of Ti(NMe2)2(NHMe2)(3-dpm3’5'CF3) (22) from single crystal X-ray diffraction ................................................................. 87 Use of 1,3-diaxial interactions to affect dipyrrolylmethane isomerizations barriers ............................................................................. 106 ORTEP diagram of the X-ray diffraction model for Ti(NMe2)2(tmcpm) (25) ........................................................................................................... 110 1H spectra of 24 at room temperature and —40 °C in CDC13 ................... 1 11 Eyring plot of pyrrolyl exchange in TiCNMe2)2(tmcpm) (25) ................. 112 Figure 6.1 Figure 6.2 Figure 6.3 Figure 6.4 Figure Al.l Figure A1.2 Figure Al.3 Figure A1.4 Figure Al.5 Figure A1.6 CS) ORTEP structure of U(NBut)2(dpmm (26) from single crystal X-ray diffraction ................................................................................................. 127 ORTEP structure of UCNBut)2(dpmmcs)(dmpe) (27) from single crystal X-ray diffraction ...................................................................................... 128 ORTEP structure of U(NBu‘)2(dpm)(THF)2 (28) from single crystal X-ray diffraction ...................................................................................... 131 ORTEP structure of U(NBut)2(pyrrmes)(THF)2(I) (29) from single crystal X-ray diffraction ........................................................................... 133 Representative plot of [l-phenylpropyne] versus time with 10 mol% Ti(NMe2)2(dpmmeS) (12) at 75 °C .......................................................... 142 Representative plot of [l-phenylpropyne] versus time with 10 mol% Ti(NMe2)2(dprn3’5‘CF3) (13) at 75 °c .................................................... 143 Representative plot of [l-phenylpropyne] versus time with 10 mol% Ti(NMe2)2(pyrr3’5’CF3)2 (14) at 75 °c ................................................... 144 Representative plot of [l-phenylpropyne] versus time with 10 mol% Ti(NMe2)2(pyrrmeS)2 (15) at 75 °c ......................................................... 145 Representative plot of [l-phenylpropyne] versus time with 10 mol% Ti(NMe2)2(pyrr4’CF3)2 (16) at 75 °c ..................................................... 146 Representative plot of [1-phenylpropyne] versus time with 10 mol% Ti(NMe2)2(pyrr‘°‘)2 (17) at 75 °c ........................................................... 147 Figure Bl.l Representative plot of [l-phenylpropyne] versus time with 10 mol% Figure Bl.2 Figure Bl.3 Figure C1.l Ti(NMe2)2(_pyrrmes)2 (15) at 100 °c ....................................................... 143 Representative plot of [l-phenylpropyne] versus time with 10 mol% Ti(indenyl)2(Me)2 (21) at 100 °C ............................................................ 149 Representative plot of [1- henylpropyne] versus time with 10 mol% I’ Ti(NMe)3(pyr2'CF3'4C6F ) (20) at 100 °C .............................................. 150 Representative plot of L1 -phenylpropyne] versus time with 10 mol% Ti(NMe)2(3-dpm3’5_c 3) (22) at 75 °c .................................................. 151 xi Figure Cl.2 Figure Cl.3 Figure D1.l Figure Dl.2 Figure D1.3 Figure Dl.4 Figure Dl.5 Figure D1.6 Figure Dl.7 Figure Dl.8 Representative plot of [l-phenylpropyne] versus time with 10 mol% Ti(NMe)2(3-dme6F3) (23) at 75 °c ....................................................... 152 Representative plot of % conversion versus time with 10 mol% Ti(NMe)2(3-dpm3’5-CF3) (22) at 75 °C using reaction calorimetry ....... 153 1 . mes H spectrum for T1(NMe2)2(dpm )(12) .............................................. 154 13C spectrum for Ti(NMe2)2(dpmmcs) (12) ............................................ 155 1H spectrum for U(NBu’)2(dpmmeS) (26) ................................................ 156 1 t mes H spectrum for U(NBu )2(dpm )(dmpe) (27) .................................... 157 31 t mes P spectrum for U(NBu )2(dpm )(dmpe) (27) ................................... 158 'H spectrum for U(NBu’)2(THF)2(dpm) (28) ......................................... 159 13c spectrum for Ti(NMe2)2(NHMe2)(pyrr3’5_CF3)2 (19) in CDC13 ....l60 5 'H spectrum for Ti(NMe2)2(NHMe2)pyrr3’ 4'33» (19) in CDC13 ....... 161 xii Boc Butbipy Cod dap szpm GC/F ID THF DME dmpe CS prrrm prn} ,S—C F 3 prrr4—CF3 prlTtol szpmmes szpm3,5—CF3 3-dpm3’5—CF3 3-dme6F3 3_prn_3,5—CF 3 LIST OF ABBREVIATIONS tert-butyloxycarbonyl 4,4’-di-tert-butyl-2,2’-bipyridine 1,5- cyclooctadiene 2-((dimethyamino)methyl)pyrrolyl 5,S-dimethyldipyrrolylmethane Gas Chromotography Flame Ionization Detector tetrahydrofuran 1 ,2-dimethoxyethane bis(dimethylphosphino)ethane 2-(mesityl)pyrrole 2-(3,5-bis(tn'fluoromethyl)phenyl)pyrrole 2-(4-(trifluoromethyl)phenyl)pyrrole 2-p-toly1-1H-pyrrole 5,5'-(propane-2,2-diyl)bis(2-mesityl-pyrrole) 5,5'-(propane-2,2-diyl)bis(2-(3,5- bis(trifluoromethyl)phenyl)pyrrole) 5,5'-(propane-2,2-diyl)bis(3-(3,5- bis(trifluoromethyl)phenyl)pyrrole) 5,5'-(propane-2,2-diy1)bis(3 -mesity1-pyrrole) 3-(3,5-bis(t1ifluoromethyl)phenyl)pyrrole xiii C6F5 3-prrr 3-(perfluorophenyl)pyrrole prrr 2-023 -4-C6F5 2-(3 ,5-bis(trifluoromethyl)phenyl)-4-(perfluorophenyl)pyrrole Hchm I ,1 -bis(a-pyrrolyl)cyclohexane Hztmcpm 1 , 1 -bis(a-pyrrolyl)-3 ,3 ,5 ,5-tetramethylcyclohexane TFA trifluoroacetic acid RT room temperature EtOAc ethyl acetate xiv Scheme 1.1 Scheme 1.2 Scheme 1.3 Scheme 1.4 Scheme 1.5 Scheme 1.6 Scheme 1.7 Scheme 2.1 Scheme 2.2 Scheme 2.3 Scheme 2.4 Scheme 2.5 Scheme 2.6 Scheme 2.7 Scheme 3.1 Scheme 3.2 LIST OF SCHEMES Comparison between condensation and titanium-catalyzed hydroamination methodologies ................................................................... 1 Bergman mechanism for hydroamination ................................................... 4 Hydroamination results with Ti(NMez)4 (3) as catalyst ............................. 5 Hydroamination/hydrosilylation by Ti(NMe2)2(Cp-SiMe2-NBut) (6) ...... 6 Synthesis of Hgdpma and Ti(NMe2)2(dpma) (8) ....................................... 9 Synthesis of szpm and Ti(NMe2)2(dpm) (9) ......................................... 11 Proposed 3-component coupling reaction mechanism involving a primary amine, alkyne, and isonitrile ....................................................... 16 Synthesis of szpm and Ti(NMe2)2(dpm (9) .......................................... 22 Proposed mechanism for titanium-catalyzed alkyne hydroamination ...... 24 Chauvin mechanism for olefin metathesis ................................................ 24 Synthesis of szpmmes, szpm3’5_CF3, Ti(NMe2)2(dpmmes) (12), and Ti(NMe2)2(dpm3’5'CF3) (13) .................................................................... 27 Synthesis of Ti(NMe2)2(pyrrm"s)2 (15) and . 3,5-CF3 T1(NMe2)2(pyrr )2 (l4) .................................................................. 30 Reaction conditions for kinetic studies ..................................................... 33 Synthesis of Ti(NMe2)2(pyrr4'CF3)2 (l6) and Ti(NMe2)2(pyrrt°l)2 (17) .......................................................................... 37 Smith borylation followed by Suzuki coupling to prepare 3-substituted pyrroles ..................................................................................................... 60 Synthesis of 3-prrr3’5-CF3 and 3-prrrC6F5 ........................................ 61 Scheme 3.3 Scheme 3.4 Scheme 3.5 Scheme 4.1 Scheme 4.2 Scheme 4.3 Scheme 5.1 Scheme 5.2 Scheme 6.1 Scheme 6.2 C6F5 Possible pyrrolyl crossover in the reaction of prrr and Ti(NMez)4 (3) ........................................................................................... 64 Synthesis of 2-(3,5-bis(trifluoromethyl)phenyl)-4- . 2-CF3-4C6F5 (perfluorophenyl)pyrrole and T1(NMe2)3(pyr ) (20) ............. 66 Conditions for kinetic study ...................................................................... 67 Synthesis of 3-aryl pyrroles ...................................................................... 83 Synthesis of 3-H2dpm3’5_CF3, 3-H2dme6F3, Ti(NMe2)2(NHMe2)(dpm3’5'CF3) (22) and Ti(NMe2)2(NHMe2)(dme6F3) (23) ........................................................ 85 Reaction conditions for kinetic study ....................................................... 88 Synthesis of szpm and Ti(NMe2)2(dpm) (9) ........................................ 105 Synthesis of Hchm, Hztmcpm, Ti(NMez)2(cpm) (24) and Ti(NMe2)2(tmcpm) (24) .......................................................................... 107 Exchange of imido ligand in U(NBut)2(THF)2I2 and U(NBu’)2(0PPh3)212 123 uuuuuuuuuuuuuuuuuuuuuuuuuuuuuuuu Synthesis of U(NBu’)2(THF)2(I)(pyrrmes) (29) ...................................... 132 xvi CHAPTER 1 Titanium pyrrolyl catalysts in C—N bond formation 1.1 Introduction The synthesis of C—N bonds is a significant process in organic chemistry. While a typical method for their preparation is the condensation of an amine or hydrazine with a ketone or aldehyde, titanium-catalyzed hydroamination is an attractive alternative when the desired product is not readily prepared using current carbonyl methodologies. While both methods are effective in preparing Simple irnine products, titanium-catalyzed hydroamination offers great synthetic utility in the selective generation of a variety of nitrogen containing products. For example, the selective syntheis of a,B-unsaturated imines from the reaction of an (LB-unsaturated ketone with a primary amine can be problematic. Competing Michael addition reactions can lead to a mixture of products which can make isolation difficult. However, titanium-catalyzed hydroamination of a 1,3-enyne with a primary amine yields the desired product selectively without unwanted by-products (Scheme 1.1). 0 NR RHN O RHN - NR + NH2R + + R2 R2 R2 R2 R R R1 1 1 ‘l . NR _ [TI] __ R, + NHZR = Hz 92 ll Scheme 1.1 Comparison between condensation and titanium-catalyzed hydroamination methodologies. In addition to the synthesis of or,B-unsaturated imines, new methodologies in the synthesis of pyrroles, hydrazones, pyrimidines, and a,B-unsaturated B-iminoamines have been developed which use titanium-catalyzed hydroamination (vide infia). The development of titanium catalysts to carry out these transformations often use the hydroamination of alkynes with primary amines as a method for evaluating the efficacy of catalyst design. While there are a plethora of titanium catalysts known to hydroaminate alkynes with primary amines, the development of catalysts through ancillary ligand studies to improve the substrate scope and enhance catalyst activity is an area of intense research. This chapter takes a cursory look at ancillary ligands for titanium catalysts with the main focus being on the enhancement of titanium pyrrolyl complexes towards the development of C—N bond forming reactions. 1.2 Hydroamination with titanium Cp complexes Cyclopentadienyl (Cp) ligands are some of the most heavily studied ancilliaries for transition metals. Fittingly, some of the first reported examples of hydroamination with titanium catalysts were ligated with Cp-based ligands.l Titanium-catalyzed hydroamination is believed to operate through a similar mechanism as the zirconocene-based system elucidated by Bergman and co-workers (Scheme 1.2).2 A primary amine is a restriction of the Bergman mechanism for hydroamination and it is assumed that all the titanium catalysts in this chapter operate via this mechanism. The first step in the catalytic cycle is the generation of a reactive metal imido species (A) which is formed from protonylsis of the precatalyst M—R bond. Introduction of an alkyne coordinates to A to yield complex B which undergoes a [2 + 2]- cylcoaddition to yield a azarnetallacyclobutene intermediate (C). Coordination of free primary amine to C in solution gives complex D which undergoes prontonlysis of the M— C bond to give bis(amido) metal complex E. Further protonation from the amide regenerates the active catalyst A and yields the imine product. Chart 1.1 Cp based titanium catalysts Ti Me ,Tif—‘NAr ( )2 (1) Doye (2) Bergman3 Doye4 NAr R L M n R [1" ( )2 R=alkyl,amido l l 122:; Fl R Ar R / N.Ar Ill ' H—>MIL1 ML n ArHN’ H" E R 3 ,l. l :81 13 ,Ar I I N MlLln | I 13 ==" Mil-In NHZAI' NHzAl‘ R D C Scheme 1.2 Bergman mechanism for hydroamination The groups of Doye and Bergman have pioneered Cp-based titanium catalysts, which have been shown to be fairly general catalysts for alkyne hydroamination (Chart 1.1).4 Bergman and co-workers revealed that TiszMez undergoes cyclopentadienyl/amide ligand exchange, which enhances the reactivity towards alkyne hydroamination. 1.3 Hydroamination with titanium amido complexes Since the first reports of hydroamination with titanium Cp catalysts, many groups have explored the use of amido ancillary ligands in hopes of generating more reactive catalysts and expanding the substrate scope. Odom reported that commercially available Ti(NMe2)4 (3) was a general hydroamination catalyst for aryl primary amines and alkynes (Scheme 1.3).5 Scheme 1.3 Hydroamination results with Ti(NMe2)4 (3) as catalyst. 10% Ti(NMez)4 (3) _ NPh 75 °C, toluene R, “2 NH2Ph + R, : R, R1, R2 = (Bun, H), (Et, El), 87 - 92 % (Ph,H) Shortly after that report, Bergman and co-workers reported the hydroamination of amino allenes and amino alkynes with sulfonamido ancillary ligands on titanium (Chart 1.2).‘5 It is also worthy to note, that the ligation of these chelating sulfonamido ligands on zirconium are relatively good asymmetric intramolecular alkene hydroamination catalysts with good yields and moderate ee’s.7 Chart 1.2 Titanium sulfonamido complexes 9 so2 so2 N N \ 0’ \TI(NM92)2 0’ TI(NM92)2 .. / -., / Iv bl! so, 802 (4) (5) Doye and co-workers reported a chelated Cp/amido titanium complex capable of hydroaminating an alkyne, followed by direct reduction via hydrosilylation with the same titanium precatalyst (Scheme 1.4).8 This synthetic procedure allows for the production of secondary amines in a one-pot procedure, without having to use stoichmetric amounts of conventional reducing agents like NaBHgCN or LiAlH4. Scheme 1.4 Hydroamination/hydrosilylation by Ti(NMe2)2(Cp-SiMe2-NBut) (6). \ \Si\ /TI(NM62)2 N ””2” /|\ (6) “I IT'l N’ I + ; )l\/R _ n2 3 PhSiHs methanol piperdine 1.4 Hydroamination with titanium amidate complexes The Schafer group has prepared a class of amidate titanium complexes capable of intermolecular hydroamination (Figure 1.1).9 The development of electron deficient amidate ligands on titanium resulted in decreased catalytic performance and unexpected ligand reactivity.lo There is supporting evidence that a more Lewis acidic metal center may facility greater hydroamination reactivity (vide infia). Figure 1.1 Titanium amidate complex 0 . (Ph—<< #TKNEtzlz N 2 (7) The strategy of using fluorine groups on the phenyl group rendered the arene susceptible to nucleophilic attack by primary amines at elevated temperatures. This resulted in the formation of HF during catalysis, which decomposed the catalyst (Equation 1.1). This observation shows there should be a judicious choice of electron- vvithdrawing substituents when selecting an ancillary ligand for titanium-catalyzed hydroamination. i i . “9:9 “1:9 F F NHZR + fl Pri A > E, Pri + HF (1.1) F F F F F F 1.5 Hydroamination with titanium pyrrolyl complexes Using deprotonated pyrroles as ancillary ligands is nothing new in transition metal chemistry. However, pyrrolyl ligands have seen increased attention as ancillaries in the field of catalysis. Among the many fields of catalysis, titanium catalyzed C—N and C— C bond forming reactions have probably been the largest application of pyrrolyl ligands. 12d Unlike many alkoxide, amide, or Cp ligands, pyrrole is not a strongly n—donating ligand. The nitrogen lone pair is delocalized around the ring to maintain aromaticity, which directly competes with rt-donation to the metal center (The aromatic stabilization energy for pyrrole is ~21 kcal/mmol).11 This decreased it-donation to the metal center results in a more Lewis acidic metal center that has proven useful in a variety of catalytic systems.12 One can easily draw analogies between the Bergman mechanism for hydroamination and the Chauvin mechanism for olefin metathesis. Both mechanisms include metal-ligand multiple bond intermediates and [2+2] cycloaddition processes with unsaturated substrates. One of the results from Schrock’s olefin metathesis studies was that the reactivity increased in his (I0 molybdenum catalysts as the metal center was made more Lewis acidic.13 Applying the same principle towards titanium catalysts for hydroamination using pyrrole-based ancillary ligands could be fruitful given the similarities in mechanisms and that pyrroles are weak rc—donors. The increased Lewis acidity of the metal could result in stronger metal-alkyne binding, facilitating the [2+2] cycloaddition. In addition, the Bronsted acidity of a coordinated amine increases upon coordination to a metal center, which may speed up the rate-limiting protonolysis step. '4 The first reported example of a pyrrolyl-based titanium catalyst for hydroamination was Ti(NMe2)2(dpma) (8), where dpma is N,N-di(pyrrolyl-a-methyl)-N— methylamine.15 szpma is readily prepared by reaction of 2 equivalents of pyrrole, 2 equivalents of formaldehyde, and an equivalent of methyl ammonium hydrochloride generating the ligand in good yield. szpma can be placed on titanium by transamination with Ti(NMe2)4 (3) to yield precatalyst 8 in excellent yield (Scheme 1.5). Scheme 1.5 Synthesis of szpma and Ti(NMe2)2(dpma) (8). H .0 + / NH HN \ O > / N \ 2 EtOH/HZO, 55 °C, 4 h l H’ILH 73% szpma + MeNH2°HC| 95% Ti(NMe2), (3) A ‘ NMBZ /| N Ti— NMe2 / ”(Mt / Me Ti(NMez)2(dpma) (8) The hydroamination of aniline and l-hexyne is carried out in 6 h at 75 °C with 8 at 10 mol% catalyst loading.l6 Complex 8 is a relatively good catalyst for the hYdroamination of aryl and alkyl amines with alkynes (Table 1.1). While most of the reactions could be carried out at 75 °C with moderate reaction times, higher temperatures were required for more difficult substrates. Table 1.1 Representative hydroamination results with 10 mol% Ti(NMe2)2(dpma) (8). Selectivity Time (h) at % Yield at (M: anti-M) Amine Alkyne 75°C [130°C] 75°C [130°C] at 75 °C [130 °C] PhNH2 Bu" _—_-_- H 6 90 50:1 Ph 2 H 8 41 3.6:1 Ph 2 Me 144[24] 99 [96] 1:24 [1:19] Et 2 Et 72 63 Ph 2 Ph 72 [74] 31[99] CyNHz Ph : H 20 50 1:6 Ph : M9 95 I29] trace [99] 1:4 Et—-=-_:——Et 72 [24] 3 [57] Ph : Ph 72 [241 0[701 The dpma ligand architecture bears pyrrolyl ligands with an r] l,r]‘-coordination in the solid state and in solution when bound to titanium. To date, this is the only bonding hapticity observed for the dpma ligand. Even though Ti(NMe2)2(dpma) (8) was a relatively good hydroamination catalyst, alterations in ligand design to create a more Lewis acidic metal were explored. The most logical modification to the dpma ligand was removal of the donor amine in the dpma backbone, which lead to the use of dipyrrolylmethanes as ligands. A variety of dipyrrolylmethanes can be prepared by reacting pyrrole with aldehydes or ketones in the presence of a catalytic amount of a Lewis or Branstead acid. '7 The neat reaction of pyrrole and acetone with a catalytic amount of trifluoroacetic acid (TFA) yields 5,5-dimethyldipyrrolylmethane, szpm. Reacting szpm with Ti(NMe2)4 (3) affords Ti(NMe2)2(dpm) (9) (Scheme 1.6). 10 Scheme 1.6 Synthesis of szpm and Ti(NMe2)2(dpm) (9). H No 25 IN 0 10%TFA N [,1 Ti(NMe2)4(3) \ D + \f T \ I l / -2HNM82 : / N/TKNMGZ)? szpm TI(NM92)2(de) (9) The solid-state structure of 9 has the pyrrolyls 111,715 -bound; however, in solution the 1H NMR spectrum shows equivalent pyrrolyls, indicative of fast pyrroyl exchange on the NMR timescale. Using line shape analysis and spin saturation transfer experiments, our group has been able to place a barrier for pyrrolyl exchange at ~10 kcal/mol."”19 Ti(NMe2)2(dpm) (9) catalyzes the reaction of aniline and l-hexyne famously in ~5 minutes with a modest 5 mol% catalyst loading. The reaction of aniline and l-hexyne is so rapid and exothermic that the reaction vessel is hot to the touch and refluxes the reagents, which results in the low yield for Entry 1 in Table 1.2.19 11 Table 1.2 Representative hydroamination results with Ti(NMe2)2(dpm) (9) as the catalyst. Selectivity Amine Alkyne Conditions °/. Yield (M: anti-M) PhNHz Bu : H 5%, 25 °C,5min 57 4011 Ph : H 5%, 25 °C,5min 41 3.6:1 Ph : Me 5%, 50 °C,6h 83 50:1 Et—z—Et 5%, 50 °C, 24 h 94 Ph : Ph 5%, 75 °C, 24h 34 CvNHz Ph : H 5%, 25 °C,1o min 54 1.621 Ph : Me 5%, 75°C, 24h 93 11:1 Et : Et 10%, 75 °C, 48h 73 ph : ph 10°/o,100°C,48h 72 The results in Table 1.2 show that 9 is a good catalyst for the hydroamination of alkyl and aryl amines with alkynes and much more reactive than Ti(NMe2)2(dpma) (8). A kinetic study comparing the various ligand architectures of 8, 9, and Cp-derived pyrrolyl complex (10), shows that the dipyrrolylmethane framework is quite a bit faster relative to the other pyrrole-based ligand frameworks (Table 1.3). 12 Table 1.3 Rate constants for Ti(NMe2)2(dpma) (8), Ti(NMe2)z(dpm) (9), Ti(NMe2)2(pyrr-Cp) (10) for the hydroamination of aniline and l-phenylpropyne. Me 0.. . 10 NH2Ph + / 10/ catalyst (0 05 M) ; NPh Ph Ph toluene, 75 °C 5 M 0.5 M d[1 -phenylpropyne] = kobst dt Entry Precatalyst kobs x 10'6 s" a NM62 1 éii—NMez (a) 11m / ”X l . / "'N‘Me N0 2 \ /T|(NM92)2 (9) [157] / N I. 10 1 3 /TI(NM82)2 ( I a Values in brackets are with chlorobenzene as solvent. Complex 9 was about 20 times faster than Ti(NMe2)2(dpma) (8) and about a 100 times faster than 10. It was proposed that Ti(NMe2)2(dpm) (9) can more readily access the nlml-isomer than 10, accounting for the increase in catalysis rate. Expanding the scope of hydroamination to include 1,3-enyne substrates allows for the synthesis of cult-unsaturated imines, which can be difficult to prepare from a purely organic synthesis approach of reacting a primary amine with an (LB-unsaturated ketone. 13 Using titanium-catalyzed hydroamination avoids the competing Michael addition products which can occur in the latter procedure. \ /TI(NM82)2 CY\N ___ | NHQCY + Qfi-_— > (1 .2) 100°C,PhCL5ll 73%: Complex 9 can conveniently hydroaminate a 1,3-enyne in 5 h at 100 °C (Equation 1.2).20 Catalyst 9 showed good reactivity with less reactive alkyl amines; however, due to the rapidity of hydroamination, 9 was problematic with sensitive substrates where potential side reactions could occur (i.e. terminal alkynes). Ti(NMe2)2(dap)2 (11) where dap is a-(dimethylaminomethyl)pyrrole proved to be the optimal catalyst for highly reactive substrates (Table 1.4). Catalyst 11 is prepared by reacting 2 equivalents of Hdap with Ti(NMe2)4 (3) (Equation 1.3).21 TMez \ Nlu \\ H NMe2 ZTiQMez 2 N + Ti(NM32)4 950/ 7: me I NMeg (1'3) \ l (3) ° N \ / Hdap , TI(NM92)2(daP)2 (11) 14 Table 1.4 Respresentative hydroamination results with Ti(NMe2)2(dap)2 (11) at 10 mol% catalyst loading. Amine 1,3-enyne Conditions % Yield Product NPh PM”? O—z 50 °C 16 h 88 O_< NPh >.—_— 50 °C, 44 n 64 >—:‘:—Ph 130 °C, 19 h 70 \ NP“ Ph NC CvNHz 0‘: 50 °C, 24 h 78 W y NCy >—_—: 50 °C, 43 h 73 The ability to make new C—N bonds using hydroamination is a desirable alternative when the condensation of carbonyls with amines is problematic. The development of well-defined pyrrolyl-based catalysts provides the organic chemist with another synthetic tool to prepare a variety of imine products. In addition to imine a products, new methodologies have been established for preparing pyrroles,22 indoles,9 and hydrazones9 via titanium-catalyzed reactions. 15 1.6 Iminoamination Development of the hydroamination reaction to yield more elaborate imine products offers more potential in preparing valuable C—N containing products with broader applications. To expand on the scope of products obtained from the hydroamination mechanism, our group developed a new 3-component coupling reaction involving the coupling of a primary amine, alkyne, and isonitrile in a single synthetic step catalyzed by a titanium pyrrolyl complex. The result is the formal irninoamination of an alkyne, which is not readily achievable using current carbonyl methodologies. The mechanism for the coupling is shown in Scheme 1.7. Scheme 1.7 Proposed 3-component coupling reaction mechanism involving a primary amine, alkyne, and isonitrile. NR NHR3 R1 my R a, i. \, [Ti] 2 NHZR K \ R R1 ~ R R '1‘ \ R ‘N 1 III] 2 I I \ [TI] R N~R3 2 cans, Ti(NMe2)2(dpma) (8) catalyzes the coupling effectively allowing for the production of a,B-unsaturated B-iminoamines (Table 1.5).23 One of the two identifiable by-products is an imine product fiom hydroamination of the alkyne by the primary amine. The other by-product is an N,N-disubstituted-formamidine, which results from the coupling of the primary amine with isonitrile. Altogether, the by-products are typically 16 found in less than 15% yield. The regioselectivities of the 3-component coupling reaction with catalyst 8 are similar to the reported regioselectivies of the hydroamination reaction.5 It is worthy to note that the coupling does not take place in the absence of the catalyst and even the by- products are not observed. Also, simple treatment of the hydroamination product with isonitrile in the presence of the catalyst results in no formation of the 3-component coupling product, therefore the azametallacyclobutene must be present to result in the formation of desired coupling product. Table 1.5 Results of 3-component coupling with Ti(NMe2)2(dpma) (8) at 10 mol% catalyst loading at 100 °C in toluene. Amine Alkyne lsonitrile % Yield PIOdUCt PhNH2 anne- czN-Bu‘ 77 Bu'HNMNPh Me '— Ph csN—Bu‘ 72 CyNH2 Bu" (1.2 : 1) Catalyst 9 is also a very good 3-component coupling catalyst and works with a variety of alkynes and aniline derivatives. Moreover, our group has expanded the applications of the 3-component coupling reaction and developed new syntheses for quinolines, pyrimidines, and pyrazoles from the 3-component coupling product.24 17 1.7 Conclusion The field of titanium-catalyzed hydroamination has seen an explosion of research since the first reported example of a titanium catalyst carrying out the hydroamination of an alkyne and a primary amine.16 The development of well-defined pyrrolyl-based titanium catalysts has led to new methodologies in ON bond forming reactions. Because so many of the applications listed are based on the hydroamination catalytic cycle, the hydroamination of alkynes can be used as a method for evaluating the efficacy of new catalyst designs. The following chapters discuss the investigation of elaborating the pyrrolyl framework in order to produce more reactive hydroamination catalysts with potential applications in multi-component coupling reactions. 18 1.8 References 1. Walsh, P. J.; Baranger, A. M.; Bergman, R. G. J. Am. Chem. Soc. 1992, 114, 1708. 2. (a)Baranger, A.M.; Walsh, P. J.; Bergman, R. G. J. Am. Chem. Soc. 1993, 115, 2753. (b) Walsh, P. J.; Baranger, A. M.; Bergman, R. G. J. Am. Chem. Soc. 1992, 114, 1708. (c) Duncan, A. P.; Bergman, R. G. Chem. Rec. 2002, 2, 431. ((1) Le, S. Y.; Bergman, R. G. Tetrahedron 1995, 51, 4255. (e) Polse, J. L.; Andersen, R A.; Bergman, R. G. J. Am. Chem. Soc. 1998, 120, 13405. 3. Johnson, J. S.; Bergman, R.G. J. Am. Chem. Soc., 2001, 123, 2923. 4. Heutling, A.; Pohlki, F .; Doye, S. Chem. Eur. J. 2004, 10, 3059. 5. Shi, Y.; Ciszewski, J. T.; Odom, A. L. Organometallics, 2001, 20, 3967. 6. Ackerman, L.; Bergman, R. G. Loy, R. N. J. Am. Chem. Soc. 2003, 125, 11956. 8. Watson, D. A. ; Chieu, M.; Bergman, R. G. Organometallics, 2006, 25, 4731. 8. Heutling, A.; Frauke Pohlki, F.; Igor Bytschkov, I.; Doye, S. Angew. Chem. Int. Ed. 2005, 44, 2951. 9. (a) Zhang, Z.; Schafer, L. L. Org. Lett, 2003, 5, 4733. (b) Li, C.; Thomson, R. K.; Gillon, B.; Patrick, B. O.; Schafer, L. L. Chem. Commun. 2003, 2462 (0) Thomson, R. K.; Zahariev, F. B; Zhang, Z.; Patrick, B. 0.; Wang, Y.A.; Schafer, L. L. Inorg. Chem, 2005, 44, 8680. 10. Bexrud, J. A.; Li. C.; Schafer, L. L. Organometallics, 2007, 26, 6366. 11. March, J. Advanced Organic Chemistry, Wiley-Interscience, New York, 4th ed., 1996, 45. 12. (a) Saeed, 1.; Katao, S.; Nomura, K. Organometallics, 2009, 28, 111. (b) Hock A. S.; Schrock, R. R.; Hoveyda, A. H. J. Am. Chem. Soc., 2006, 128, 16373. (c) Mazet, C.; Gade, L. H. Organometallics, 2001, 20, 4144. (d) Odom, A. L. Dalton T rans., 2005, 225 and references therein. (e) Schouteeten, S.; Allen, 0. R.; Haley, A. D. ;. Ong, G. L.; Jones, G. D.; Vicic, D. A. J. Organ. Chem. 2006, 4975. 13. Schrock, R. R.; Murdzek, J. S.; Bazan, G. C.; Robbins, J.; DiMare, M.; O’Regan, M. J. Am. Chem. Soc. 1990, 112, 3875. 14. Fulton, J. R.; Sklenak, S.; Bouwkamp, M. W.; Bergman, R. G. J. Am. Chem. Soc. 2002, 124, 4722. 15. Harris, S.A.; Ciszewski, J. T.; Odom, A.L. Inorg. Chem, 2001, 40, 1987. 19 16. Cao, C.; Ciszewski, J. T.; Odom, A.L. Organometallics, 2001, 20, 5011. 17. (a) Littler, B. J .; Miller, M. A.; Hung, C. H.; Wagner, R. W.; O’Shea, D. F.; Boyle, P. D.; Lindsey, J. S. J. Org. Chem. 1999, 64, 1391. (b) Patchanita T. P.; Bhise, A. D.; Taniguchi, M.; Lindsey J. S. J. Org. Chem, 2006, 71, 903. 18. Shi, Y.; Hall, C.; Ciszewski, J. T.; Cao, C.; Odom, A. L. Chem. Commun. 2003, 586. 19. Swartz, D.L.; Odom, Dalton Trans. 2008, 4254. 20. Cao, C.; Li, Y.; Shi, Y.; Odom, A. L. Chem. Comm. 2004, 2002. 21. For Ti(NMe2)2(dap)2 as a hydrohydrazination catalyst see (a) Cao, C. Shi, Y.; Odom, A.L. Org. Lett, 2002, 4, 2853. (b) Banerjee, S.; Shi, Y.; Cao, C.; Odom, A.L. J. Organo. Chem. 2005, 5066. 22. Ramanathan, B.; Keith, A. J.; D. Armstrong D.; Odom, A.L. Org. Lett, 2004, 6, 2957. 23. Cao, C.; Shi, Y.; Odom, A. L. J. Am. Chem. Soc. 2002, 125, 2880. 24. Majumder, S.: Odom, A. L.; manuscripts in preparation. 20 CHAPTER 2 Synthesis, structure, and hydroamination kinetics of 2,2’-diaryldipyrrolylmethane and bis(Z-arylpyrrolyl)titanium complexes 2.1 Introduction The generation of new carbon-nitrogen bonds is a significant process in organic chemistry. The most atom economical process for the generation of amines and imines is through hydroamination, which is the formal addition of an N—H bond across an unsaturated C—C bond. Many natural products and pharmaceutical drugs contain C—N bonds; therefore these heteroatom molecules offer pharmaceutical applications as well as other industrial applications in the way of dyes, detergents, and fungicides.l Pyrrolyl- ligated titanium complexes provide very reactive catalysts for hydroamination. This chapter discusses 2,2’-diaryldipyrrolylmethane and bis(2-arylpyrrolyl)titanium complexes as competent hydroamination catalysts and gauges the efficacy of their design through kinetic studies. 21 2.2 Titanium Pyrrolyl Hydroamination Titanium-catalyzed intermolecular hydroamination has been widely studied and has led to new methodologies in the synthesis of imines,2 hydrazones,3 indoles,2 5 tautomers of 1,3-dimines,6 and tautomers of 1,3- pyrroles,4 unsaturated imines, iminohydrazones.7 In addition, a variety of nitrogen containing heterocycles have been synthesized by intramolecular cyclization.8 There are many ancillary ligands on titanium known to mediate these types of transformations. The ligands employed include 0 l alkoxides,9 amides,1 amidates,l and pyrroles.12 Pyrrolyl-ligated titanium complexes provide very reactive catalysts capable of extremely fast hydroamination. Perhaps the most active precatalyst known for simple alkyne hydroamination as of 2008 was a dipyrrolylmethane-ligated titanium complex Ti(NMe2)2(dpm) (9), where dpm is 5,5-dimethyldipyrrolylme1hane. The dipyrrolylmethane ligand is readily prepared from acetone and pyrrole in the presence of trifluoroacetic acid (TFA). The ligand can then be placed on titanium by transamination with Ti(NMe2)4 (3) to yield the precatalyst in good yields (Scheme 2.1).l3 Scheme 2.1. Synthesis of szpm and Ti(NMe2)2(dpm) (9). H HN \ 5 ‘ o H ___. ' \ min, / 53% / 90.,3 _ N (9) szpm TilNM92)2(de) (9) The use of pyrrole-based ancillary ligands offers several advantages compared to other architectures. First, pyrroles can be easily manipulated into multidenate ligands by a standard set of condensation reactions that take advantage of the nucleophilic nature of the pyrrole ring. Therefore, several classes of ligands can be synthesized in a small 22 number of steps (e.g. Mannich reaction) with a vast degree of steric and electronic variance. Secondly, pyrroles are relatively weak 1t donors compared to their alkoxide or amide counterparts due to the nitrogen lone pair delocalization by its participation in the ring’s aromaticity. This decreased donation to titanium results in a more Lewis acidic metal center. Titanium-catalyzed hydroamination is believed to operate through a smiliar mechanism as the Bergman mechanism for hydroamination elucidated by Bergman and co-workers using a zirconocene-based system (Scheme 2.2).14 The first step in the catalytic cycle is the generation of a reactive titanium imido species (A). Introduction of an alkyne coordinates to A which undergoes a [2 + 2]-cylcoaddition to yield a azametallacyclobutene intermediate (B). Coordination of free primary amine to B in undergoes prontonlysis of the Ti—C bond to give bis(amido) titanium complex C. Further protonation from the amide regenerates the active catalyst A and yields the imine product. Applying the steady-state approximation to the intermediates in the mechanism shown in Scheme 2.2, the following rate law can be derived (Equation 2.1). k k k alk ne catal st v= 1231 V II V] (2.1) k-,(k2+k3 [amine]) 23 Scheme 2.2 Proposed mechanism for titanium-catalyzed alkyne hydroamination NAr Ft, R2 LnTl(NHAl')2 u NH2Ar k1 k1 R /1 ’Ar er a N l Fl A 2 k4 k-2 k2 Rt R, ,Ar N R2 / N’Ar )E'l’iLn ,TiLn F12 ATHN B C W NH2Ar One can easily draw connections between the Group-4 Hydroamination Mechanism for alkyne hydroamination and the Chauvin mechanism for olefin metathesis (Scheme 2.3). Scheme 2.3. Chauvin mechanism for olefin metathesis R R M=CH2 R \ ___. 7/ F \K R M “[1 D I R R G Fl /—_— \§\‘ M=/ R H Both mechanisms are known to include metal-ligand multiple bond intermediates (A and F). Another key step is the [2 + 2]-cyclization between the C—C unsaturated bond 24 and the metal ligand multiple bond (B and G). Detailed olefin metathesis studies done by Schrock and co-workers found that increasing the Lewis acidity of d0 metal centers led to an increase in catalysis rates.15 Given the similarities in mechanism and the known improvement in catalyst activity in (10 Schrock carbenes bearing electron deficient alkoxides on Lewis acidic metal centers, it is expected that decreasing the donor ability of the pyrrole ligands could have a positive effect on catalyst reactivity. This chapter describes the effects on structure and catalysis 2-aryl substituents have on the dpm framework. Complexes containing 2-aryl bis(pyrrolyl) ligands were also synthesized to evaluate the effect the linker in the dpm framework has on structure and catalysis. 25 2.3 Hydroamination kinetics with titanium dipyrrolylmethane derivatives The selective generation of 2-arylpyrroles can be achieved by the seminal methodology established by Sadighi and co-workers (Equation 2.1).16 This technique allows for the production of 2-arylpyrroles on multigram scales. For this study, I wanted to investigate pyrroles of varying electronic and steric profiles. RZP RP [PdI/ Pit—C cat. )1 U + Zl'lClg + Ar-X - MAr (2.1) THF, 60 - 100 °C ’ For this study, I prepared 2-Ar-pyrroles where Ar = 4-(CF3)C6H4, 4-(CH3)C6H4, 3,5-(CF3)2C6H3, and 2,4,6-(CH3)3C6H2. The typical procedure for the synthesis of dipyrrolyhnethanes involves the use of a large excess of pyrrole with respect to the aldehyde or ketone (Scheme 2.1). However, in the preparation of 2-substituted dpm derivatives, pyrrole is the limiting reactant with excess acetone as the electrophile. The reaction is smoothly catalyzed by TFA to produce the ligands in good yield. The initial 5,5-dimethyldipyrrolylmethane derivatives prepared were 2,9-[3,5-(CF3)2C6H3]-5,5- 3,5-CF3 dimethyldipyrrolylmethane (szpm ) and 2,9-[2,4,6-(CH3)3C6H2]—5,5- dimethyldipyrrolylmethane (szpmmes). These complexes were then reacted with Ti(NMe2)4 (3) to yield the corresponding dipyrrolylmethane metal complexes, 3,5—CF3 Ti(NMe2)2(dpmmes) (12) and Ti(NMe2)2(dpm ) (13) (Scheme 2.4). 26 Scheme 2.4 Synthesis of szpmmes, szpm3’5‘cp3, Ti(NMe2)2(dpmmes) (12), and TiCNMe2)2(dpm3’5'CF3) (13). \21 0 10°/ TFA O H "N \ + 25 /U\ 1h, RT. neat t I N 68% / H 2dpmmes . Ti(NMe2)., (3) 3° 4 0512, so °c 18 n - 2 l-iNMc2 Ti(NM°2)2(dem°s) (12) 27 CF3 H N o 10% TFA + 25 = W /u\ 3 h, RT, neat 72°/o CF3 92% 0E12. T Ti(NM62)2(dpm3'5‘CF3) (13) Single crystal X-ray diffraction shows that the pyrrole rings are in a 111,115 conformation in the solid state for both Ti(NMez)2(dpmmes) (12) and Ti(NMe2)2(dpm3’5 CF3) (13). The crystal structure of 13 is shown in Figure 2.1. This is similar to previous studies carried out by our group and the Love group on Ti(NMe2)2(dpm) (1).17 The pyrrole rings in Ti(NMe2)2(dpm) (9) are in an minis-conformation in the solid state as well. In cold solutions on the NMR timescale, resonances consistent with the solid state structure are observed. As the solution warms, resonances for the til-pyrrolyl and 115- pyrrolyl coalesce, and the fast exchange limit is reached well before room temperature. The barrier for pyrrolyl ligand exchange was measured at 10 kcal/mol using line shape analysis.4 It is believed that the pyrrolyl exchange occurs through a 111,111-isomer.18 The 28 magnitude for the barrier of pyrrolyl ligand exchange and the mechanism for isomerization are consistent with other known pyrrolyl isomerizations in the literature.19 F(l3l) _ 2Q -... F(153) N2 '- F(l5|) 5‘.) a, ‘3 . T‘ ‘5?" I v/y _ F(152) N13) 0.4 a t. - F(232) F(233)I' ?? a,” 1‘ .: F(252) F031) I-d '4‘ F(251) H253) Figure 2.1 ORTEP structure from single-crystal X-ray diffraction of Ti(NMe2)2(dpm3’5'CF3) (13). Selected bond distances (A) and angles (deg): Ti—N(3) 1.875(5), Ti-N(4) 1.892(5), Ti-N(2) 2.048(5), Ti-N(1) 2.400(6); N(4)-Ti-N(3) 107.1(2), N(4)-Ti-N(2) 102.6(2), N(3)-Ti-N(2) 104.8(2). Detailed NMR studies were carried out on 12 and 13 to investigate pyrrolyl exchange. Consistent with previous results, 12 and 13 Show resonances for equivalent pyrrolyls at room temperature, indicative of rapid n',n5-isomerization. Cooling the solutions to —60 °C showed no difference in the 1H NMR spectrum. Parkin and Tanski have reported that increased sterics lower the barrier for pyrrolyl exchange.18 These results are consistent with their findings, and assuming a pyrrolyl isomerization is taking place, one can place a maximum on the pyrrolyl exchange barrier of 5 kcal/mol. 29 To investigate the effect of removing the methylene linker in the dpm architecture on catalysis, I synthesized two bis(pyrrolyl) derivatives bearing the same aryl 3,5-CF3 substituents. Ti(NMe2)2(pyrr ) (14) and Ti(NMe2)2(pyrrmes)2 (15) were prepared by 3,5-CF3 reacting two equivalents of prrrmes and prrr with Ti(NMe2)4 (3) to give the corresponding metal complexes (Scheme 2.5). The solid-state structures for 14 and 15 are shown in Figures 2.2 and 2.3. Both structures show both pyrroles in an nlznl-binding mode. Cooling solutions to —60 °C in the NMR probe showed no new resonances in the baseline. Scheme 2.5 Synthesis of Ti(NMe2)2(pyrrmeS)2 (15) and Ti(NMe2)2(pyrr3’5'CF3)2 (14). F330 CFS H CF3 N 2 I/ + Ti(NMe2)4 r 0512. RT. 6h (MeaNiaTi N \ CF 3 38% 3 -2HNM62 — 2 Ti(NM62)2(pyr3'5'CF3)2 (14) H 2 N . I / + TI(NM62)4 T 3 74% - 2 HNMe2 OEt2, 60 °C, 40 h (MeszTitN \ ) 2 TI(NM92)2(PYrm°s)2 (15) 30 Figure 2.2 ORTEP structure from single-crystal X-ray diffraction on TiCNMez)2(pyrrm°s)2 (15). Selected bond distance (A) and angles (deg): Ti-N(3) 1.844(4), Ti-N(4) = 1.865(4), Ti-N(2) 1.971(4), Ti-N(1) 2.007(4); N(3)-Ti-N(4) 109.3(2), N(3)-Ti- N(2) 107.5(2), N(4)-Ti-N(4) 109.3(2), N(3)-Ti-N(1) 112.0(2), N(4)-Ti-N(1) 114.1(2), N(2)-Ti-N(1) 105.2(2). With these complexes in hand, I set out to answer several questions. First, how would sterics in the 2-position affect the catalysis rate? Second, how does the linker affect the catalysis rate? Third, can it be shown experimentally that electron-withdrawing substituents increase the catalytic activity as proposed in the Introduction? To test the kinetic viability of these complexes, I chose a standard set of reaction conditions (Scheme 2.6). The reactions were run pseudo-first order in aniline. Aniline was chosen because it has a large catalyst scope and runs at a reasonable rate compared to other amines. The limiting reagent was l-phenylpropyne. This alkyne was chosen for a couple of reasons. 31 First, it has good regioselectivity with most catalysts and runs at a reasonable rate. Second, an internal alkyne was required due to the rapidity of hydroamination with terminal alkynes with catalysts like Ti(NMe2)2(dpm) (9). In addition to testing the kinetics of the titanium dpm derivatives, I also wanted to compare Ti(NMe2)4 (3), which serves as a very reasonable hydroamination catalyst as well, but is generally limited in substrate scope to aryl amines and can oligomerize terminal alkynes. F(252) F(253) 1., .. 1 _ F(25 1’) F(23l) _ . . ,_ r “-1 .v ”1- 12) '. F 233 F(232) 3’ ( )- f) ‘41 é "‘3, F053) ‘. ”1 F052) "' s , x vi F(l5l) ,9, F032) F(l31) .‘5 F(l33) ‘ Figure 2.3 ORTEP structure from single-crystal X-ray diffraction on Ti(NMe2)2(pyrr3’5'CF3)2 (14) Selected bond distances (A) and angles (deg): Ti-N(1) 1.989(4), Ti-N(2) 2.011(4), Ti-N(3) 1.855(4), Ti-N(4) 1.848(4), N(4)-Ti-N(3) 108.88(17), N(4)-Ti-N(l) 114.72(16), N(3)-Ti-N(1) 107.32(l6), N(4)-Ti-N(2) 107.97(16), N(3)-Ti- N(2) 1100.69(16), N(1)-Ti-N(2) 107.3605). 32 These complexes were evaluated based on Ti(NMe2)2(dpm) (9). Complex 9 catalyzes the reaction to completion efficiently in less than 4 hours at 75 °C with a rate constant of 1976 i 130 ><10'7 s". Often, catalytic reactions Show a first-order dependence in catalyst concentration, which is true in this case. Changes in catalyst concentration track linearly with the rate constant (Figure 2.4). Scheme 2.6 Reaction conditions for kinetic studies Me °. . 10 NH2Ph + / 10 / catalyst 0 05 M > NPh Ph Ph toluene-d8, 75 °C 5 M 0.5 M d[1 -phenylpropyne] dt = kobst Figure 2.4 Plot of catalyst concentration of 9 versus pseudo first-order rate constant for kinetic reaction conditions. 2000 . . l 1800. 16WI 14W'r 1200 1 km x 10'7 S'1 1000» —y III £11.73 '0 51848)! R8 0.99951 l 1 A 0.02 0.025 0.03 0.035 0.04 0.045 0.05 0.055 [91 M 33 While most of the catalysts were quite regioselective, monitoring the reaction by disappearance of l-phenylpropyne was preferential due to possible rate inconsistencies by measuring the formation of products which may include the other regioisomer. The comparison of these catalysts by relative reaction rate was measured by kobs. A representative plot of the disappearance of ln[l-phenylpropyne] versus time with TiCNMe2)2(dpm) (9) as the catalyst is shown in Figure 2.5. Figure 2.5 Representative plot of ln[l-phenylpropyne] vs time with complex 9 as the hydroamination catalyst. 05 - —y = 0.56599 - 0.21074x R: 0.99946 .1 _ E -1.5 . E >5 . _ E 2 .fl 9‘ 3' -2.5 _ .3 ._ .35 I I r l l l 1 0 2 4 6 8 1O 12 14 Time X 103 (s). The results of this kinetic study are shown in Table 2.1 for catalysts 9 and 12 - 15. The errors are based on 99% confidence level, with at least three repeated runs. The average error in rate constants was ~10% and varied from as little as 4% to as much as 20%. 34 Table 2.1 Observed rate constants for catalysts Ti(NMe2)2(dpm) (9), TiINMethdpmm“) (12). TiINMe2)2(dpm3’5‘CF3) (13). TitNMenztpyt-P’S‘CBh (14), Ti(NMe2)2(pyrrmes)z (15). Entry Catalysta kobs (x 10-7 3")" 1 Ti(NMe2)2(dpm) 1976 :1: 130 9 2 Ti(NMe2)2(dpm3’5_CF3) 780 a 30 13 3 Ti(NMe2)2(dpmmes) 403 a: 80 12 4 Ti(NMe2)2(pyrr3'5‘CF3)2 1275 a 72 14 5 TiCNMe2)2(pyn-mes) 769 a: 30 15 3 Conditions are shown in Scheme 2.6. bAll errors are at the 99% confidence limit with at least three repeated runs. 35 It is quite clear to see from Table 2.1 that substitution on the 2-position of the dpm architecture results in a significant decrease in reaction rate. This decrease in rate can be attributed to increased sterics of the dpm framework near the substrate binding site. Comparatively, the unlinked bis(pyrrolyl) complexes still had a slower rate than Ti(NMe2)2(dpm) (9), but showed a faster rate than their linked dpm derivaties by ~1.5 times. A possible explanation for the increase in hydroamination reaction rate for the unlinked pyrrolyl catalysts is their ability to rotate their bulky substiuents away from the substrate binding site unlike the dpm derivatives, leading to a more open metal center. While each of the dpm derivatives and bis(pyrrolyl) catalysts bearing an electron withdrawing substituent resulted in faster rates, these examples do not allow the separation of steric and electronic factors. As a result, two additional catalysts were prepared Ti(NMe2)2(py1-r4'CF3 ) (16) and Ti(NMez)2(pyrrt°l)2 (17) that differ only in the donor ability of the substituent in the 4- position of the aromatic group. The pyrroles, 2-(4-(trifluoromethyl)phenyl)pyrrole (prrr4'CF3) and 2-(4-tolyl)pyrrole (l-lpyrrtOI), were synthesized using Sadighi’s protocol (Equation 2.1). The pyrroles were then placed on titanium by transamination with Ti(NMez)4 (3) to give the corresponding metal complexes (Scheme 2.7). 36 Scheme 2.7 Synthesis of Ti(NMe2)2(pyrr4'CF3)2 (16) and Ti(NMe2)2(pyrr‘°')2 (17). CF3 CF + Ti(NM62)4 = a 0512, 25°C. 36 h. (MegN)2Ti + (MezNthVCN \ > 0606' 75 0C : (MGZN)2T1< (2.2) 2 — 2 F C N \ 3 \ CF, While there is no definitive answer as to whether the mono(pyrrolyl) or bis(pyrrolyl) or a mixture of both are carrying out the catalysis, it can be said that the active species is not Ti(NMe2)4 (3) or a species not containing the pyrrolyl ligand. Throughout all the catalyses no free prrr resonances were observed in the 1H NMR spectrum, therefore the maximum amount of Ti(NMe2)4 (3) that can be generated is half the amount of the concentration in our kinetic studies. Because the rate scales linearly with catalyst concentration, full disproportionation to Ti(NMe2)4 (3) and Ti(pyrr)4 would 39 give a rate constant of 433 X 10'7 5‘, half the Ti(NMe2)4 (3) rate under our standard conditions. Since all of the catalysts are faster than this, it can be said definitively that Ti(NMez)4 (3) is not the active species. 40 2.4 Conclusion Using readily available 2-substituted pyrroles, a route to 2,9- diaryldipyrrolylmethanes has been developed where the synthesized pyrroles can be used as the limiting reagent in condensation with acetone. Placing these new ligands on titanium is readily accomplished by reaction with Ti(NMe2)4 (3), and the resulting complexes show haul-coordination of the dipyrrolylmethane in the solid state. However, the barrier for pyrrolyl exchange in these substituted dipyrrolylmethanes is quite low and is below what could be measured by variable temperature 1H NMR spectroscopy. Catalysis with these new dpm complexes was slower than with unsubstituted 9, which are assigned to steric inhibition by the bulky groups near the substrate-binding site. Consistent with this, simple bis(pyrrolyl) complexes without the methylene linker and with the same substituents show faster catalysis rates, which is likely due to free rotation of the steric bulk away from the substrate binding site in the bis(pyrrolyl) derivatives. Using these bis(pyrrolyl) catalysts, I were able to show experimentally that hydroamination activity can be increased by adding electron-withdrawing groups to pyrrolyl substituents. Comparison of various catalyst architectures shows that the pyrrolyl catalysts are quite rapid. It should be noted that this does not necessarily imply that these catalysts are superior for any particular application. The “best” catalyst for any particular application is a function of availability, selectivity, and activity with a particular set of substrates. For example, the fastest catalyst studied here, 9, is a poor catalyst for some applications, e.g. hydrohydrazination,20 and the only catalyst studied that leads to any product for others, e.g. the synthesis of 1-phenyl-2,5-dibenzylpyrrolyl from 1,6-diphenyl-l,5-hexadiyne and 41 aniline.8 The aim is to optimize the activity of the most promising catalysts for these reactions, which my current results suggest are the dipyrrolylmethane complexes of titanium, especially Ti(NMe2)2(dpm) (9). From these experiments, I am discovering what electronic features and steric profiles encourage these useful C—N bond forming reactions. 42 2.4 Experimental General Considerations. All manipulations of air sensitive compounds were carried out in an MBraun drybox under a purified nitrogen atmosphere. Anhydrous ether was purchased from Columbus Chemical Industries Inc., and pentane and toluene was purchased fiom Spectrum Chemical Mfg. Corp., were purified by sparging with dry N2, then water was removed by running through activated alumina systems purchased from Solv—Tek. Hexanes and ethyl acetate were purchased from Mallinckodt-Baker Inc., and reagent grade acetone was purchased from Fisher Scientific and distilled from CaSO.t under N2 and stored over 4A molecular sieves. Trifluoroacetic acid was purchased from Aldrich and used as received. Aniline was purchased from Matheson, Coleman and Bell Mfg. and was distilled twice from calcium hydride under vacuum. l-phenylpropyne was purchased from GFS Chemical, vacuum distilled, and then passed over two columns of neutral alumina. Ti(NMe2)4 (3)21 was prepared using the literature procedures. 2-(2,4,6- t1imethylphenyl)-lH-pyrrole (prrrmes) and 2-[3,5-bis(trifluoromethyl)phenyl]-1H- 3,5-CF3 pyrrole (prrr ) were synthesized according to literature methods.23 Deuterated solvents were dried over purple sodium benzophenone ketyl (C6D6) or phosphoric anhydride (CDCl3) and distilled under a nitrogen atmosphere. Deuterated toluene 7 was dried by passing it over two columns of neutral alumina. 1H and '3C spectra were recorded on Inova-300 or VXR-500 spectrometers. All spectra were referenced internally to residual protiosolvent ('H) or solvent (BC) resonances. 1H and 13C assignments were confirmed when necessary using two-dimensional lH——'H and lH—BC correlation NMR experiments. Chemical shifts are reported in ppm and coupling constants reported in Hz. 43 General Considerations for X-Ray Diffraction. Crystals grown from concentrated solutions at —35 °C quickly were moved from a scintillation vial to a microscope slide containing Paratone N. Samples were selected and mounted on a glass fiber in wax and Paratone. The data collections were carried out at a sample temperature of 173 K on a Bruker AXS platform three—circle goniometer with a CCD detector. The data were processed and reduced utilizing the program SAINTPLUS supplied by Bruker AXS. The structures were solved by direct methods (SHELXTL v5.1, Bruker AXS) in conjunction with standard difference Fourier techniques. General Procedure for Kinetics. All manipulations were done in an inert atmosphere drybox. In a 2 mL volumetric flask was loaded the catalyst (10 mol %, 0.1 mmol), aniline (0.931 g, 911 ,uL, 10 mmol), l—phenylpropyne (0.116 g, 125 ”L, 1 mmol), and ferrocene (0.056 g, 0.3 mmol) as an internal standard. The solution was then diluted to 2 mL with deuterated toluene. An ample amount of solution (~0.75 mL) was put into a threaded J. Young NMR tube that was sealed with a cap and then wrapped with Teflon tape. The tube was then removed from the drybox and heated at 75 °C in the NMR spectrometer. The relative l-phenylpropyne versus ferrocene concentration was monitored as a function of time. The fits are to the exponential decay of the starting material using the scientific graphing programs Origin or KaleidaGraph. The exact expression used to fit the data was Y, = YOo + (Y0 - Yoo) exp-“b5! where Y = [1- phenylpropyne] at time = t (Yr), infinity (Yoo), or initial (Yo). The variables Yoo, Y0, and kobs were optimized in the fits.22 44 Synthesis of 2-(4-methylphenyl)-lH-pyrrole (prrrml) (D Q/ Z’:F” 0_cr Under an atmosphere of dry nitrogen a threaded Schlenk tube was loaded with sodium pyrrole (4.69 g, 52.6 mmol), Zan (7.17 g, 52.6 mmol) and a stirbar. To that same vessel was added of THF (40 mL) slowly (caution: exothermic). After 10 min, Pd(OAc)2 (20 mg, 0.5 mol%) and 2-(di-tert-butylphosphino)biphenyl (26 mg, 0.5 mol%) were added to the Schlenk tube. The tube was then capped, taken from the dry box, and connected to a Schlenk line. Under a continuous flow of nitrogen the screw cap was removed, and 4-bromotoluene quickly was added. The screw cap was replaced after addition. The headspace in the Schlenk tube was evacuated and then placed in a 100 °C oil bath for 24 h. After 24 h, the tube was removed from the oil bath and allowed to cool to room temperature. The cap was removed, and the solution was transferred to a separatory funnel. The tube was rinsed with OH; (30 mL) and of H20 (30 mL), which were then added to the separatory funnel. The aqueous layer was extracted with OH; (3 X 50 mL); the combined organic layers were collected and dried over MgSO4. The solution was then filtered and concentrated in vacuo. Purification of the crude product was accomplished by column chromatography on silica gel using an eluting solution of hexaneszethyl acetate (9:1) to yield an white solid (1.44 g, 54%). mp. 145 °C. 1H NMR (500 MHz, CDC13): 8.2 (br s, 1 H, H“), 7.37 (d, Jun = 8.46 Hz, 2 H, H3 or H“), 7.18 (d, JHH = 8.07 Hz, 2 H, H8 or H“), 6.82 (m,1 H, H°, Hb or H“), 6.50 (m, 1 H, H“, Hb or H“), 6.31 (m, 1 H, H“, H“, or H“), 2.39 (s, 3 H, H). 13c {‘H}NMR (125 MHz,CDC13): 135.85 (ci or c“), 132.19 (ci or C“), 129.97 (0‘), 129.50 (0“ or C“), 123.77 (ch or c“), 118.42 45 (C“, C“ or C“), 109.90 (C“, Cc or C“), 105.31 (C°, C“, or C“), 21.38 (C’). Elemental Analysis (Experimental) Calc. C: (84.10) 84.04, H: (7.02) 7.05, N:(8.66) 8.91. 46 F3 Synthesis of 2-(4-trifluoromethylphenyl)-lH-pyrrole (prrrw ) Ha FC. 31 \ '0 d Under an atmosphere of dry nitrogen, a threaded Schlenk tube was loaded with sodium pyrrole (3.6 g, 40.4 mmol), ZnClz (5.5 g, 40.4 mmol) and a stirbar. To that vessel was added 32 mL of THF slowly (caution: exothermic). Afier 10 min, Pd(OAc)2 (15 mg, 0.5 mol%) and 2-(dicyclohexylphosphino)biphenyl (24 mg, 0.5 mol%) were added to the Schlenk tube. The screwcap was replaced, and the tube was removed from the dry box and connected to a Schlenk line. Under a continuous flow of nitrogen the screwcap was removed and 4-bromobenzotrifluoride (3 g, 13 mmol) quickly was added. The screwcap was replaced, and the headspace in the tube was evacuated. The tube was then placed in a 80 °C oil bath where it was allowed to react for 20 h. After the reaction was complete, the tube was removed from the oil bath and allowed to cool to room temperature. The cap was removed, and the solution was transferred to a separatory funnel. The tube was rinsed OH; (30 mL) and H20 (30 mL), which were then added to the separatory funnel. The aqueous layer was extracted with OEtz (3 X 50 mL); the combined organic layers were collected and dried over MgSO4. The solution was then filtered and concentrated in vacuo. Purification of the crude product was accomplished by column chromatography on silica gel using an eluting solution of hexaneszethyl acetate (9:1) to yield a white solid (2.53 g, 90%). mp. 158 °C. 1H NMR (500 MHz, CDC13): 8.46 (br s, l H, H“), 7.59 (d, Jun = 7.73 Hz, 2 H, H“), 7.53 (d, Jun = 7.73 Hz, 2 H, H“), 6.91 (m, 1 H, H“: H“, or H“), 6.62 (m, 1 H, H“, H°, or H“), 6.33 (m, 1 H, H“, H“, or H“). 13C {‘H} NMR (125 MHz, CDC13): 135.91 (Cf), 130.6 (C'), 127.85 (q, JCF = 33 Hz, C“), 125.92 (q, JCF = 33 Hz, C“), 47 124.29 (q, Jcp = 272 Hz, C“), 123.58 (C8) , 120.1 (C“, C“, or C“), 110.65 (C“, C“, or C“), 107.7 (Cc, Cd, or Cb). Elemental Analysis (Experimental) Calc. C: (62.22) 62.56, H: (3.54) 3.82, N: (6.51) 6.67. Caution! The lithium pyrrolide of 2-(4-(trifluoromethyl)pheny1)pyrrole (prrr4- CF3) was recently prepared in our laboratory by addition of n-butyllithium to prrr4- CF3, which allowed apparent production of the desired lithium salt. This compound, Li(pyrrCF3), was found to be explosive in the solid state under an inert atmosphere. It is likely that o-CFg-aryl pyrrolides can undergo a similar decomposition. Consequently, extreme caution should be used if alkali-metal salts are produced of aryl-substituted pyrroles containing CF 3 groups in the ortho or para positions of the arene.23 It is unknown if the same decomposition can occur in an explosive manner in solution as well, but we also urge caution with solutions of such compounds. We have produced the lithium salts of pyrroles containing m-CF3-aryl groups on numerous occasions, and, thus far, these have not undergone the same explosive decomposition. 48 Synthesis of 2,9-bis[3,5-bis(trifluoromethyl)phenyl]-5,5-dimethyldipyrrolylmethane (szpms’wm) A l-neck 14/20 25 mL round-bottom flask was charged with 2-[3,5- bis(trifluoromethyl)phenyl]-lH-pyrrole (0.4 g, 1.4 mmol) and acetone (2.08 g, 36 mmol). The flask then was sealed with a septum. The solution was stirred at room temperature while degassed under a flow of argon. After 15 min, trifluoroacetic acid (0.4 g, 3.6 mmol) was added via syringe. The solution was allowed to stir for 3 h under an argon atmosphere. The reaction was quenched with ~15 mL of 0.1 M NaOH solution. The resulting mixture was transferred to a separatory funnel where the aqueous layer was extracted with EtZO (2 x 15 mL). The combined organic layers were dried over MgSO4 and filtered. The solvent was removed in vacuo to give a purple oil. The oil was then tritrated with pentane to yield a pink solid (0.311 g, 72%). mp. 130 °C. 1H NMR (500 MHz, CDC13): 8.11 (br s, 2 H, H“), 7.73 (s, 4 H, H‘), 7.6 (s, 2 H, H'), 6.6 (m, 2 H, H“ or H“), 6.2 (m, 2H, H“ or H“), 1.76 (s, 6 H, H“). 13C {‘H}NMR (125 MHz, CDCl3): 141.64 (C“ or C“), 134.4 (C“ or C“), 132.18 (q, Jcp = 31.94 Hz, C’), 129.03 (C“), 123.28 (q, Jcp = 272 Hz, C"), 123.15 (q, Jcp = 2.54 Hz, C‘), 119.1 (q, JCF = 3.91 Hz, C'), 108.64 (C“ or C‘), 107.09 (C° or Cf), 35.83 (C’), 29.09 (Cd). Elemental Analysis (Experimental) Calc. C: (54.63) 54.19, H: (3.14) 3.03, N: (4.52) 4.68. 49 Synthesis of 2,9-bis(2,4,6,-trimethylphenyl)-5,5-dimethyldipyrrolylmethane CS (szpmm ) A l-neck 14/20 25 mL round-bottom flask was charged with 2-(2,4,6— trimethylphenyD-lH-pyrrole (0.25 g, 1.3 mmol) and acetone (1.96 g, 33 mmol) then sealed with a septum. The solution was stirred at room temperature while degassed under a flow of argon. After 15 min, trifluoroacetic acid (0.384 g, 3.3 mmol) was added via syringe. The solution was allowed to stir for 1 h. The reaction was quenched with ~15 mL of 0.1 M NaOH solution. The resulting mixture was transferred to a separatory funnel where the aqueous layer was extracted with B20 (2 X 15 mL). The combined organic layers were dried over MgSO4 and filtered. The solvent was removed in vacuo to give an orange oil. The oil was then tritrated with pentane to yield an orange solid. (0.187 g, 68%). mp. 115 °C. 1H NMR (500 MHz, CDC13): 7.57 (br s, 2 H, H‘), 6.89 (s, 4 H, H"), 6.09 (app t, JHH = 2.87 Hz, 2 H, H“ or H‘), 5.88 (app t, JHH = 5.86 Hz, 2 H, H“ or H‘), 2.2 (s, 6 H, H'“), 2.0 (s, 12 H, H“ ), 1.6 (s, 6 H, H“). 13C {‘H}NMR (125 MHz, CDCl;): 138.54 (C', C“, or C“), 138.31 (C', C“, or C“), 137.38 (C', C“, or C“), 130.83 (C“ or C“), 128.76, 127.98 (C‘), 107.5 (Ce or Cf), 103.4(C° or C’), 35.30 (C°), 28.95 (C’), 20.99 (Cm), 20.52 (C’). Elemental Analysis (Experimental) Cale: C, (84.85) 84.83, H: (8.37) 8.35, N: (6.78) 6.82. 50 Synthesis of Ti(NMe2)z(dpmmes) (12): Under an atmosphere of dry nitrogen, a threaded pressure tube was loaded with Ti(NMe2)4 (3) (0.316 g, 1.4 mmol), szpmmes (0.578 g, 1.4 mmol), and E120 (8 mL). The pressure tube was sealed with a Teflon screw cap, taken out of the dry box, and put in a 50 °C oil bath, where it was lefl to react for 18 h. The pressure tube was then removed from the oil bath and taken back into an atmosphere of dry nitrogen, where the volatiles were removed in vacuo to yield an orange oily solid. Crystallization from pentane yielded an orange solid (0.225 g, 30%). mp. 155 °C (dec). 1H NMR (500 MHz, CDC13): 6.80 (s, 4 H, H"), 6.37 (d, Jun = 2.80 Hz, 2 H, Hf or H’), 6.28 (d, JHH = 2.77 Hz, 2 H, Hf or H°), 2.59 (s, 12 H, H“), 2.23 (s, 6 H, H”), 2.18 (s, 12 H, H“), 1.94 (s, 6H, H“). l3C {‘H} NMR (125 MHz, CDCl3) 159.15 (C“), 138.38 (Ci or C'), 138.26 (Ci or C'), 136.27 (C“ or C8), 132.54 (C8 or C“), 128.52 (C‘), 113.70 (C“ or C“), 107.01 (C“ or Cf), 47.45 (C'), 39.26 (C’), 30.15 (Cm), 21.67 (C’), 20.92 (C°). After many attempts at elemental analysis, satisfactory results were not obtained. 1H and 13 C NMR spectra are included in the Appendix to demonstrate purity. 51 Synthesis of Ti(NMe2)z(dpm3’5'CF3) (13): Under an atmosphere of dry nitrogen a vial was loaded with Ti(NMe2)4 (3) (0.131 g, 0.584 mmol) in EtzO (3 mL). In a 20 mL scintillation vial was loaded 2,9-bis[3,5- bis(trifluoromethyl)phenyl]-5,5-dimethyldipyrrolylmethane (0.350 g, 0.584 mmol) in OEtz (3 mL). The solutions were put into a cold well where they sat until nearly frozen. 3'5”“ was added to the cold solution of 3. The solution To a thawing solution of szpm was stirred at room temperature for 4 h. Volatiles were removed in vacuo to yield an orange solid (0.395 g, 92%). mp. 150 °C (dec). 1H NMR (300 MHz, CDC13): 7.96 (s, 4 H, H‘), 7.63 (s, 2 H, H'), 6.55 (d, JHH = 3.19 Hz, 2 H), 6.44 (d, Jun = 3.18 Hz, 2 H), 2.6 (s, 12 H, H“), 1.8 (s, 6 H, H“). ”C {'H} NMR (125 MHz, CDC13): 162.84 (C“), 138.18 (C8 or C8), 138.04 (C“ or C8), 131.17 (q, Jcp = 33.6 Hz, C“), 126.26 (q, C“), 119.55 (q, Jcp = 3.7 Hz, C'), 110.22 (CC or C‘), 109.53 (C“ or C‘), 45.88 (Cm), 39.67 (0529.33 (C“). Elemental Analysis (Experimental) Calc. C: (50.45) 50.84, H: (3.75) 3.85, N: (7.48) 7.65. 52 Synthesis of Ti(NMe2)2(pyrrmes)z (15): k (MGZN)2 \ T I N / C Under an atmosphere of dry nitrogen a threaded pressure tube was loaded with Ti(NMe2)4 (3) (0.075 g, 0.334 mmol), prrimes (0.124 g, 0.667 mmol), and EtzO (5 mL). The pressure tube was then sealed with a Teflon screwcap and wrapped with Teflon tape. The pressure tube was then taken out of the dry box and put into a 60 °C oil bath for 40 h. Afier 40 h, the pressure tube was taken back into an atmosphere of dry nitrogen, and the volatiles were removed in vacuo to yield a yellow solid (0.125 g, 74%). mp. 160 °C (dec). 1H NMR (500 MHz, CDC13): 6.83 (s, 4 H, H“), 6.81 (m, 2 H, H“, H“, or H“), 6.19 (m, 2 H, H“, H“ or H0, 5.91 (m, 2 H, H“ ,H“, or H“), 2.77 (s, 12 H, H"), 2.28 (s, 6H, H“), 2.1 (s, 12 H, H8). ”C {H} NMR (125 MHz, CDC13): 139.2 (C‘), 137.54 (C‘, C“, or C“), 136.85 (C‘, C“, or C“), 134.65 (C‘, C“, or C“), 127.6 (C“), 123.09 (C', C“, or C'), 108.6 (C' C“ or C“), 108.3 (C“, C“, or C“), 44.14 (C'), 21.04 (C“), 20.36 (C8). Elemental Analysis (Experimental) Cale. C: (70.98) 71.42, H: (8.01) 7.99, N: (10.97) 11.10. 53 Synthesis of Ti(NMe2)2(pyrr3'5'C”)2 (14) l (MGZN)2\ 1'1 CF3 Nd . bc 19 SF“ 2 Under an atmosphere of dry nitrogen, a 20 mL scintillation vial was loaded with Ti(NMe2)4 (3) (0.150 g, 0.669 mmol) in OEtz (3 mL). A separate vial was loaded prrr3’5_C123 (0.372 g, 1.33 mmol) in OEtz (5 mL). prrr3’5‘CF3 was then added to the vial containing 3 at room temperature. The solution was allowed to stir at room temperature for 6 h. Crystallization from pentane yielded the bis(pyrrolyl) as an orange solid (0.175 g, 38%). m.p. 86 °C (dec). 'H NMR (500 MHz, CDCh): 7.74 (s, 4 H, H‘), 7.61 (s, 2 H, H“), 6.89 (m, 2 H, H“, H“, or H“), 6.48 (m, 2 H, H“, H“, or H“), 6.25 (m, 2 H, H“, H“, or H“), 3.0 (s, 12 H, H“). ”C {‘H} NMR (125 MHz, CDCh): 138.88 (C“), 137.70 (C“), 131.62 (q, .10: = 33 Hz, C8), 126.63 (C“, C“, or C“), 125.96 (C‘), 123.35 (JCF = 273 Hz, C“), 119.1 (C‘), 111.86 (C“, C“, or C“), 111.20 (C“, C“, or C“), 44.53 (C“). Elemental Analysis (Experimental) Calc. C: (48.48) 48.57, H: (3.53) 3.49, N: (7.78) 8.09. 54 Synthesis of Ti(NMe2)2(pyrr4'CF3)2 (l6) (MeN). 2 2;” N i I/de hcfi bc 9 Under an atmosphere of dry nitrogen, a 20 mL scintillation vial was loaded with Ti(NMe2)4 (3) (0.061 g, 0.272 mmol), prrr4_CF3 (0.115 g, 0.544 mmol), CB; (3 mL), and a stir bar. The vial was then capped and allowed to stir for 36 h at room temperature. The volatiles were removed in vacuo to yield an orange oil. Crystallization from pentane yielded an orange solid (0.069 g, 45%). mp. 72 °C. 1H NMR (500 MHz, CDCl3): 7.47 (d, Jun = 7.88 Hz, 4 H, H8), 7.45 (d, JHH = 7.89 Hz, 4 H, Hf ), 7.00 (m, 2 H, H“), 6.36 (m, 2 H, H“ , H“, or H“), 6.27 (m, 2 H, H“, H“ or H“), 3.16 (s, 12 H, H“). ”C {‘H} NMR (125 MHz, CDC13): 140.43 (C“), 139.50 (C“), 127.64 (q, Jcp = 33 Hz, C“), 126.56 (C‘), 126.51 (C“, C“, or C“), 125.12 (q, Jcp = 11.78 Hz, C8), 124.31 (q, Jcp = 272 Hz, C“), 111.07 (C“, C“, or C“), 109.86 (C“, C“, or C“), 44.68 (C“). Elemental Analysis (Experimental) Calc. C: (55.83) 56.13, H: (5.11) 4.71, N: (10.29) 10.07 55 Synthesis of Ti(NMez)2(pyrrt°l)2 (17): a (“'92le Ti . 9 N b I e 1—_§\:"c d Under an atmosphere of dry nitrogen a threaded pressure tube was loaded with 2 prrrml (0.115 g, 0.732 mmol), Ti(NMe2)4 (2) (0.082 g, 0.366 mmol), GB; (3 mL), and a stirbar. The pressure tube was then sealed with a Teflon screwcap and then wrapped with Teflon tape. The tube was removed from the dry box and placed in a 60 °C oil bath, where it was left to react for 24 h. After the reaction was complete, it was taken back into an atmosphere of dry nitrogen. The reaction mixture was transferred to a 20 mL scintillation vial where the volatiles were concentrated in vacuo to yield an orange oil. Crystallization from pentane yielded an orange solid (0.060 g, 37%). mp. 71 °C. 1H NMR (500 MHz, CDC13): 7.18 (d, Jun = 7.68 Hz, 4 H, Hh or H“), 7.04 (d, Jun = 7.77, 4 H, H“ or H8), 6.90 (m, 2 H, H“, H“, or H“), 6.27 (m, 2 H, (H“, H“, or H“), 6.18 (m, 2 H, H“ ,H“, H“), 2.99 (s, 12 H, H“), 2.29 (s, 6 H, H“). ”C {'H} NMR (125 MHz, CDCl;): 141.11 (C‘), 135.53 (C’), 134.57 (C“), 128.82 (C“ or C8), 126.97 (C“ or C8), 125.67 (C“), 110.05 (C“, C°’ or C“), 107.68 (C“ ,C“ or C“), 44.68 (C“), 21.10 (C“). Elemental Analysis (Experimental) Calc. C: (69.53) 69.94, H: (7.60) 7.19, N: (12.17) 12.49. 56 2.5 References 1. 10. 11. 12. 13. 14. 15. Chenier, P.J. Survey of Industrial Chemistry, John Wiley & Sons, New York, 1986; 82. (a) Zhang, Z.; Schafer, L. L. Org. Lett. 2003, 5, 4733. (b) Cao, C.; Ciszewski, J. T.; Odom, A. L. Organometallics, 2001, 20, 5011. (a) Cao, C.; Shi, Y.; Odom, A. L. Org. Lett. 2002, 4, 2853. (b) Khedkar, V.; Tillack, A.; Michalick, M.; Beller, M. Tetrahedron Lett. 2004, 45, 3123. (c) Ackerman, L.; Born, R. Tetrahedron Lett. 2004, 45, 9541. (d) Khedkar, V.; Tillack, A.; Michalik, M.; Beller, M. Tetrahedron 2005, 61, 7622. Cao, C.; Shi, Y.; Odom, A. L. J. Am. Chem. Soc. 2003, 125, 2880. C30, C.; Li, Y.; Odom, A. L. Chem. Commun. 2004, 2002. C30, C.; Shi, Y.; Odom, A. L. J. Am. Chem. Soc. 2003, 125, 2880. Banerjee, S.; Shi, Y.; Cao, C.; Odom, A.L. J. Organomet. Chem. 2005, 690, 5066. Majumder, S.; Odom. A. L. Organometallics, 2008, 27, 1174. Khedkar, V.; Tillack, A.; Beller, M. Org. Lett. 2003, 5, 4767. Shi, Y.; Ciszewski, J. T.; Odom, A. L. Organometallics, 2001, 20, 3967. Bexrud, J. A.; Li, C.; Schafer, L. L.Organometallics, 2007, 26, 6366. Swartz, D.L.; Odom. A.L. Organometallics, 2006, 26, 625. Littler, B.J.; Miller, M.A.; Hung, C. H.; Wagner, R.W.; O’Shea, D.F.; Boyle, P.D.; Lindsey, J. S. J. Org. Chem. 1999, 64, 1391. (a)Baranger, A.M.; Walsh, P. J.; Bergman, R. G. J. Am. Chem. Soc. 1993, 115, 2753. (b) Walsh, P. J.; Baranger, A. M.; Bergman, R. G. J. Am. Chem. Soc. 1992, 114, 1708. (c) Duncan, A.P.; Bergman, R. G. Chem. Rec. 2002, 2, 431. ((1) Le, S. Y.; Bergman, R. G. Tetrahedron 1995, 51, 4255. (e) Polse, J. L.; Andersen, R. A.; Bergman, R. G. J. Am. Chem. Soc. 1998, 120, 13405. Schrock, R. R.; Murdzek, J. S.; Bazan G. C.; Robbins, J.; DiMare, M.; O’Reagan, M. J. Am. Chem. Soc. 1990, 112, 3875. 16. Reith, R. D.; Mankad, N. P.; Calimano, E.; Sadighi, J. P. Org. Lett. 2004, 6, 3981. 17. A. Novak, A. J. Blake, C. Wilson and J. B. Love, Chem. Commun., 2002, 2796. 57 18. Shi, Y; Hall, C.; Ciszewski, J. T.; Cao, C.; Odom. A.L. Chem. Commun., 2003, 586. 19. Tanski, J. M.; Parkin, G. Organometallics 2002, 21, 587. 20. Banerjee, S.; Odom, A. L. unpublished results. 21. Bradley, D. C.; Thomas, I.M.; J. Chem. Soc., 1960, 3859. 22. Espenson, J. H. Chemical Kinetics and Reaction Mechanisms; McGraw-Hill: New York, 1995. 23. A similar caution has been issued by Wigley and co-workers concerning Li-4- CF 3C6Hs, which likely undergoes a similar decomposition: Weller, K. J .; Gray, S. D.; Briggs, P. M.; Wigley, D. E. Organometallics 1995, 14, 5588. 58 CHAPTER 3 Synthesis, structure, and hydroamination reactivity of electron deficient titanium complexes bearing bis- and mono(pyrrolyl) ligands 3.1 Introduction Pyrrole-based ancillary ligands on titanium are a useful class of compounds I 2 known to participate in catalytic hydroamination, iminoarnination, and hydrohydrazination.3 In the previous chapter, I reported that placing aryl-substituents on the 2-position of the pyrrolyl provided effective catalysts for the hydroamination of primary amines and alkynes. Using these bis(2-arylpyrrolyl) titanium catalysts, I was able. to show experimentally that hydroamination activity can be increased by adding electron- withdrawing groups to the pyrrolyl ligand, however the addition of sterically hindered groups in these positions inhibits access to the metal center negatively affecting catalysis rates.4 In an attempt to increase the hydroamination activity of these bis(pyrrolyl) titanium catalysts, moving electron-withdrawing substituents from the 2—position to the 3- position of the pyrrole may provide greater access to the binding site resulting in enhanced hydroamination activity. This chapter discusses the synthesis and hydroamination catalysis of 3-arylpyrrolyl titanium complexes. 59 3.2 Results and Discussions The synthesis of 3-arylpyrroles can be achieved by the seminal methodology established by Smith and co-workers, which begins with an iridium-catalyzed reaction of N-Boc-pyrrole with 4,4,5,5-tetramethyl-1,3,2-dioxaborolane (HBpin) to generate 3- (pinacolboryl)-N-Boc-pyrrole. The aryl group is then installed by Suzuki cross-coupling to generate 3-aryl pyrroles on multi-grarn scales (Scheme 3.1).5 Scheme 3.1 Smith borylation followed by Suzuki coupling to prepare 3- substituted pyrroles. (Cod = 1,5-cyclooctadiene, Butbipy = 4,4'-di-tert-butyl-2,2'- bipyridine, Boc = tert-butyl-carboxylate). Boc H 1.5% [Ir(OMe)(Cod)]2 Boc EN) + 1.5 O’ on 3% Butbipy : Q / N Hexane, 60 °C, 12 h 85 A: Pro HBpin O> F \ / Bpin DME, 100 °C F F 8 d F F F 53/° 170°C /° 20min Consequently, moving the aryl groups from the 2-position to the 3-position on the pyrrole reduces sterics around the metal center allowing for the retention of dimethylamine. As expected the longest Ti—N bond distance is from the dimethylamine 61 ligand, 2.248(4) A, which sits in the axial position trans to one of the pyrrolyl ligands. The average Ti—N(NMe2) bond distance is 1.867 A, while the average pyrrolyl Ti—N bond distance is 2.075 A. These bond distances are quite similar to previously reported derivatives.4 (57 ,3 Run/___} "“5, F(231) *— 2 "in; F233 5A" "L ." ( ) N“ a i x, . 1“ a". ‘ ‘ i ( "‘" 9 Ti ‘i' a. l w 3, .- F:(252) G : 11 > F033) r, ’1 F(253) 77"“ F031) F(251) “’1. 7““: - a, 1) “‘ F(152) F(151) Figure 3.1 ORTEP structure from single crystal X-ray diffraction on Ti(NMe2)2(NHMe2)(3-pyr3’5'CF3)2 (19) with thermal ellipsoids at 50% probability level. Selected bond distances (A) and angles (deg). Ti-N(1) 2.098(4), Ti-N(2) 2.053(5), Ti- N(3) 1.862(5), Ti-N(4) 1.873(5), Ti-N(5) 2.248(4), N(1)-Ti-N(2) 87.89(18), N(2)-Ti-N(3) 116.7(2), N(3)-Ti-N(4) 111.9(2), N(4)-Ti-N(5) 90.21(18), N(2)-Ti-N(4) 130.8(2), N(2)- Ti-N(5) 81.54(17), N(3)-Ti-N(5) 91.33(18), N(1)-Ti-N(3) 99.01(19), N(1)-Ti-N(4) 87.89(18), N(1)-Ti-N(5) 167.72(18). The 1H spectrum of 19 at room temperature is inconsistent with the solid-state structure showing equivalent pyrrolyls. This is indicative of a rapid isomerization on the 62 NMR timescale. Cooling solutions of 19 in CDC13 to —60 °C in the NMR probe produces a rather complex lH Spectrum. Small peaks in the aromatic region present in the room temperature lH spectrum, which were presumed to be impurities, intensify at —60 °C in the 1H spectrum and the aromatic peaks associated with the pyrrolyls of 19 split into new resonances. New peaks also appear in the methyl region and the methyl peaks associated with the dimethylamine ligand begin to split as well. Due to the complexity of the 1H spectrum at —60 °C a definitive structural assignment of the peaks in the spectrum was not possible. One possible explanation for this behavior is that there could be additional isomers of 19 in solution causing the complexity of the 1H spectrum at room temperature as well as at depressed temperatures. NHM62 CF 3 H MeZN...,,,l. SQ 2 | N + Ti(NM92)4 , M92” | 01:3 / OEtz, 25 °C, 6 h N CF3 F30 “3 F30 Ti(NMez)2(NHMez)(3-pyrr 3M“) (19) Surprisingly, reacting 2 equivalents of 3-prrrC‘SFS with Ti(NMe2)4 (3) does not seem to yield a single metal complex Ti(NMe2)2(NHMe2)(per6F5)2 or Ti(NMe2)2(NHMe2)2(per6F5)2. The 1H NMR spectrum of the orange solid obtained from the reaction mixture is rather perplexing. Integration of the peaks did not correspond to a single metal complex. Variable temperature 1H NMR experiments were carried out in toluene-d3 proved to be fairly informative. An increase in temperature in the NMR probe results in new resonances in the baseline of the 1H spectrum as well as significant shifts 63 in the methyl groups assigned to NMe2 groups. One possible explanation for this C6F5 behavior is that upon formation of Ti(NMe2)2(NHMe2)n(pyr )2, it disproportionates to yield a mixture of mono- and 1ris(pyrrolyl) complexes (Scheme 3.3) and pyrrolyl crossover is occurring at elevated temperatures. Attempts to isolate a single metal complex from the reaction of 3-prrrC6F5 and Ti(NMe2)4 (3) via crystallization were successful. Single crystals of Ti(NMe2)2(NHMe2)(pyrrC6F5)2 were grown from a OEt2/pentane solution at -—35 °C, proving that Ti(NMe2)2(NHMe2)(pyrrC6F5)2 is formed at low temperatures. Scheme 3.3 Possible pyrrolyl crossover in the reaction of prrrC‘SF5 and Ti(NMe2)4 (3). K1 4 \ / ,t 2 Ti(NM62)4 = 2 Ti(NMez)2(NHMez)n(pyr°°F5)2 F (3) F F n F F Ti(NMez)(NHMe2)n T(NM92)3(NHM92)11 '1‘ N \ / F F \ / + F F F F F F F F 3 The reaction of aniline and l-phenylpropyne under pseudo-first order reaction conditions at 75 °C with 19 as the catalyst, results in the reaction being ~10% complete after 10 h. One possible explanation for the slow catalysis may be that pyrrolyl crossover is occurring at elevated temperatures, similar to the process depicted in Scheme 3.3. This assertion warrants further investigation. 64 F(451) «1 12(453) .. 41f'f’.F(452) C. ,5 H430 ..,. f) - ’5." .- m’ _ “Q . ‘__ of", v", \; \”" F(432) T1 2‘: g; C F(433) (381141 '1'“ y) -; F(2) 4 ._..,, 1.13““ F(3) F14) Figure 3.2 ORTEP structure from Single crystal X-ray diffraction on ri(NMe2)3(pyr2'CF3‘4C“F5) (20) with thermal ellipsoids at 50% probability level. Selected bond distances (A) and angles (deg). Ti-N(1) 1.865(4), Ti-N(2) 1.861(5), Ti-N(3) 1.865(5), Ti—N(4) 2.074(5), N(1)-Ti-N(2) 106.7(2), N(2)-Ti-N(3) 108.3(2), N(3)-Ti-N(4) 1 17.09(1 9), N(2)-Ti-N(4) 1 10.2(2), N(1)-Ti-N(4) 106.35(19). In an attempt to reduce amine coordination, pyrrolyl crossover and also to increase the hydroamination activity, I prepared 2-(3,5-bis(trifluoromethyl)pheny1)-4- (perfluorophenyl)pyrrole and the corresponding titanium complex (Scheme 3.4). 65 Scheme 3.4 Synthesis of 2-(3,5-bis(1rifluoromethyl)phenyl)-4- (perfluorophenyl)pyrrole (prnlcm'cm) and Ti(NMe2)3(pyr2'C““’4C8“8) (20). F30 Boc H 1.5% [lr(OMe)(Cod)]2 OF“ N 0’8 t - 8°C 1 ‘ 3% Bu blpy a N + .5 O a F30 \ / N Hexane.60°C,18h \ / 95% Bpin CFs HBpin 100°C 1. Pd(PPh3)4. K3PO4 DME 2. areal:5 8d CF3 Boc 165°C,10min \ / 20% CSFS CFS Tl(NM62)4 (3) £12 86 % Ti(NMe2)3(pyr2-CF34C6F5) (20) Reacting 1 equivalent of prrlg'CFM‘C6F5 with Ti(NMe2)4 (3) generates Ti(NMe2)3(pyr2'CF3'4C6F5) (20) in 86% yield. A structure of 20 is shown in Figure 3.2. The synthesis of Ti(NMe2)2(pyr2'CF3'4C6F5)2 would have been preferential, but 2-CF3-4C6F 5 reacting two equivalents of prr with Ti(NMe2)4 (3) produces only the 66 2-CF3-4C6F5 mono(pyrroyl) titanium complex. Attempts to prepare Ti(NMe2)2(pyr )2 by salt 2-CF3-4C6F5 metathesis through the reaction of 2 equivalents of Lipyr with Ti(NMe2)2Cl2 yielded only the mono(pyrrolyl) complex as well. Complex 20 is a very competent hydroamination catalyst carrying out the reaction of l-phenylpropyne and aniline in less than 6 hours at 75 °C. Reintroducing sterics in the 2-position of the pyrrolyl in 20 proved advantageous in preparing a much more reactive catalyst compared to 19. Complex 20 was the first mono(pyrrolyl) titanium complex I had prepared and isolated since working with these pyrrolyl titanium complexes. I wanted to evaluate the hydroamination reactivity of 20 relative to Ti(NMe2)2(pyrrmes)2 (13), a bis(pyrrolyl) titanium catalyst from Chapter 2, as well as Ti(indenyl)2Me2 (21) which has been shown by Doye and co-workers to be a fairly general alkyne hydroamination catalyst, through kinetic studies.6 The conditions for the kinetic study are Shown in Scheme 3.5 Scheme 3.5 Conditions for kinetic study Me 10% catalyst 0.05 M NPh 10 NH2Ph + é : Ph/ toluene-d8, 100 °C )k/ Ph 5M 0.5M d[1 -pheny| propyne] dt = kobst Both 20 and Ti(NMe2)2(pyrrmes)2 (l3) carry out the reaction of aniline and 1- phenylpropyne effectively at 75 °C. While Ti(indenyl)2Me2 (21) was active at 75 °C, the results were inconsistent at this temperature, which was apparently due to a catalyst activation period. Assuming that there was an activation problem, I incubated the catalyst 67 with the aniline portion at 100 °C prior to alkyne addition. However, this still did not result in good reproducibility of the kinetics under these conditions. Consequently, I ran the reactions with Ti(indenyl)2Me2 (21) at 100 °C, which afforded plots that reliably fit to first-order kinetics. The results of the kinetic study are shown in Table 3.1. The errors are at the 99% confidence limit and based off as least three repeated runs and range from as little as 7% to as much as 27%. The results in Table 3.1 suggest that the pyrrolyl framework is quite effective relative to other catalyst architecture. Comparison of Ti(indenyl)2(Me)2 (21) with Ti(NMe2)2(pyrrmes)2 (13) under identical conditions reveals that this pyrrolyl catalyst is about a factor of 4 times faster for these substrates. Complex 20 was significantly faster than Ti(Indenyl)2(Me)2 (21), which resulted in catalysis about 8 times faster, and was twice as fast as Ti(NMe2)2(pyrrmes) (13). 68 Table 3.1 Comparison of Ti(NMe2)2(pyr2‘CF3‘4C6F5) (20), Ti(NMe2)2(pyrrmeS)2 (l3), and Ti(indenyl)2Me2 (21). Entry Catalysta kobs (x10'7 5-1) 2 (MegN)2T'1~ 2 l3 3 Ti(Indenyl)2Me2 21 7366 :1: 2002 888 m 61 3888 i 764 69 3.3 Conclusion Using the seminal C—H activation/borylation work by Smith and co-workers followed by Suzuki coupling, two new 3-substituted pyrroles were synthesized as well as 3,5—CF3 a 2,4-disubstituted pyrrole. Placing 3-prrr on titanium is readily accompliShed by reaction with Ti(NMe2)4 (3). The decreased sterics of the 3-substituted pyrrole compared to the 2-substituted derivative results in an increased coordination number with retention . . . 3,5—CF3 of HNMe2. However, subjectlng T1(NMe2)2(pyrr )2(NHMe2) (19) to the hydroamination of aniline and 1-phenylpropyne resulted in extremely slow catalysis. The reaction of 2 equivalents of 3-prrrC6F5 with Ti(NMe2)4 (3) does not yield the anticipated bis(pyrrolyl) complex. Variable temperature experiments on the orange solid from the reaction mixture suggest ligand exchange is occurring to give a mixture of titanium complexes that are relatively hydroamination inactive. While the bis(3-arylpyrrolyl) titanium complex resulted in disappointing catalysis, fast hydroamination kinetics were achieved with Ti(NMe2)3(pyrr2'CF3'4'C6F5) (20). The comparison of 20 relative to Ti(NMe2)2(pyrrmeS)2 (13) and Ti(indenyl)Me2 (21) under identical reaction conditions shows that the 20 is significantly faster than the bis(pyrrolyl) complex 13 and the Cp-based system 21. Whether the faster rates with 20 are due to electronic factors resulting from the different ligand architecture or steric constraints, or a combination of the two effects is currently unknown, which is subject of ongoing scrutiny. The further optimization of these 3-aryl pyrroles towards the synthesis of more reactive titanium catalysts is addressed in the following chapter. 70 3.4 Experimental General Considerations: All manipulations of air sensitive compounds were carried out in an MBraun drybox under a purified nitrogen atmosphere. Anhydrous ether was purchased from Columbus Chemical Industries Inc. Pentane and toluene were purchased from Spectrum Chemical Mfg. Corp. These were purified by sparging with dry N2, then dried by running through activated alumina systems purchased from Solv-Tek. Hexanes and ethyl acetate were purchased from Mallinckodt Baker Inc. Pinacolborane was purchased fi'om BASF and was used as received. N-Boc-pyrrole was vacuum distilled and stored under a purified nitrogen atmosphere over molecular sieves. Di-tert- butyl-dicarbonate (Boc anhydride) was purchased from Oakwood Chemical and was used as received. Aniline was purchased from Matheson Coleman and Bell Mfg. and was distilled twice from calcium hydride under vacuum. l-phenylpropyne was purchased from GF S Chemical, distilled under vacuum, and then passed over two columns of neutral alumina. Ti(NMe2)4,7 2-[3,5-bis(trifluoromethyl)phenyl]-1H-pyrrole (prrr3’5’ C173)? and Ti(indenyl)2(Me)210 were synthesized according to literature methods. Deuterated solvents were dried over purple sodium benzophenone ketyl (C6D6) or phosphoric anhydride (CDCl3) then distilled under a nitrogen atmosphere. Deuterated toluene was dried by passing it through two columns of neutral alumina. 1H and 13C spectra were recorded on Inova-300 or VXR-SOO spectrometers. All spectra were referenced internally to residual protiosolvent (H) or solvent (13C) resonances. General Considerations for X—Ray Diffraction. Crystals grown from concentrated solutions at —35 °C quickly were moved fi'om a scintillation vial to a microscope slide containing Paratone N. Samples were selected and mounted on a glass fiber in wax and 71 Paratone. The data collections were carried out at a sample temperature of 173 K on a Bruker AXS platform three-circle goniometer with a CCD detector. The data were processed and reduced utilizing the program SAINTPLUS supplied by Bruker AXS. The structures were solved by direct methods (SHELXTL v5.1, Bruker AXS) in conjunction with standard difference Fourier techniques. General Procedure for Kinetics. All manipulations were done in an inert atmosphere drybox. In a 2 mL volumetric flask was loaded the catalyst (10 mol %, 0.1 mmol), aniline (0.931 g, 911 14L, 10 mmol), l-phenylpropyne (0.116 g, 125 uL, 1 mmol), and ferrocene (0.056 g, 0.3 mmol) as an internal standard. The solution was then diluted to 2 mL with deuterated toluene. An ample amount of solution (~0.75 mL) was put into a threaded J. Young tube that was sealed with a cap and then wrapped with Teflon tape. The tube was then removed from the drybox and heated at 75 or 100 °C in the NMR spectrometer. The relative l-phenylpropyne versus ferrocene concentration was monitored as a function of time. The fits are to the exponential decay of the starting material using the scientific graphing programs Origin or KaleidaGraph. The exact kObst where Y = [1- expression used to fit the data was Y, = Yoo + (Yo - Yoo) exp- phenylpropyne] at time = t (Y t), infinity (Yoo), or initial (Yo).8 The variables 1’00, Y0, and kobs were optimized in the fits. 72 Synthesis of 3-(3,5-bis(trifluoromethyl)phenyl)-lH-pyrrole (3-prrr3HF3) \ZI CF3 F3C Under an atmosphere of dry N2, tetrakis(triphenylphosphine)palladium (0.817 g, 0.07 mmol) was loaded into a 20 mL scintillation vial in DME (5 mL). To the same vial was added 3,5-bis(trifluoromethyl)bromobenzene (4 g, 13.69 mmol), and the two were stirred 10 min before being transferred to a Schlenk tube. To the Schlenk tube was added 3- (pinacolboryl)-N-Boc-pyrrole (3.9 g, 13.31 mmol) in DME (10 mL). Finally, K3P04 (5.31 g, 19.97 mmol) was added. The tube was capped, taken out the dry box, and connected to a Schlenk line. The headspace of the Schlenk tube was removed, and the vessel was placed in a 100 °C oil bath for 24 h. When the reaction was complete as judged by GC-FID, the solution was transferred to a separatory funnel. The tube was rinsed with OEt2 (30 mL) and H20 (30 mL). The solution was extracted with OEt2 (2 x 50 mL). The combined organic layers were collected and dried over MgSO4. The solution was filtered and volatiles were removed in vacuo to yield a viscous oil. The oil was then transferred to a threaded Schlenk tube equipped with a stirbar. Then it was placed under a continuous flow of N2. The reaction vessel was heated in an oil bath at 170 °C with stirring for ~ 20 min. The black solution was then run through a plug of silica gel with methylene chloride (500 mL). The volatiles were removed in vacuo to yield a white solid (2.81 g, 76%). M.p. 106-109 °C lH NMR(CDC13, 500 MHz): 8.37 (br s, 1 H), 7.90 (s, 2 H), 7.65 (s, 1 H), 7.18 (m, 1 H), 6.87 (q, 1 H), 6.58 (m, l H). 13C {'H} 73 NMR (CDC13, 125 MHz): 138.04, 131.81 (q, Jcp = 31.51 Hz), 124.85 (dd), 123.77 (q, 1cF = 277.72 Hz), 122.53, 119.78, 118.77 (sept., Jcp = 4.63 Hz), 115.74, 106.63. Anal. (Found) Calcd: C, (51.63) 51.29; H, (2.53) 2.33; N, (5.02) 5.03. 74 Synthesis of 3-(perfluorophenyl)-lH-pyrrole H N l / Under an atmosphere of dry N2, tetrakis(triphenylphosphine)palladium (0.107 g, 0.092 mmol) was loaded into a 20 mL scintillation vial in DME (5 mL). To the same vial was added pentafluorobromobenzene (2.2 g, 8.9 mmol), and the two were stirred 10 min before being transferred to a Schlenk tube. To the tube was added 3-(pinacolboryl)-N- Boc-pyrrole (2.63 g, 8.98 mmol) in DME (10 mL). Finally, K3PO4 (3.58 g, 13.46 mmol) was added. The tube was capped, taken out the dry box, and connected to a Schlenk line. The headspace was removed, and the vessel was placed in a 100 °C oil bath for 8 d. When the reaction was complete as judged by GC-F ID, the tube was rinsed with OEt2 (30 mL) and H20 (30 mL). The solution was transferred to a separatory funnel and extracted with OEt2 (2 x 50 mL). The combined organic layers were collected and dried over MgSO4. The solution was then filtered and the volatiles removed in vacuo to yield a viscous oil. The oil was then transferred to a threaded Schlenk tube equippedwith a stirbar. Then it was placed under a continuous flow of N2. The reaction vessel was heated in an oil bath at 170 °C with stirring for ~ 20 min. The black solution was then run through a plug of Silica gel with methylene chloride (500 mL). The volatiles were removed in vacuo to yield a white solid (1.1 g, 53%). M.p 48-51 °C. 1H NMR (CDC13, 500 MHz): 8.46 (s, 1 H), 7.29 (s, 1 H), 6.89 (app s, 1H), 6.68 (app s, 1 H). ”C {'H} NMR (CDCl3, 125 MHz) 6 145.10-144.08 (m), l43.05—142.98 (111), 13941-13929 (m), 75 13899-13888 (m), 137.41-137.29 (m), 137.18-36.93 (m), 119.48 (t, Jcp = 6.81 Hz), 118.43, 111.23-110.01 (m), 109.34 (t, Jcp = 5.98 Hz). Anal. (Found) Calcd: C, (51.37) 51.52; H, (1.72) 1.73; N, (5.85) 6.01. 76 Synthesis of Ti(NMezh(3-pyrr3’5'CF3)2(NI-1Me2) (19) NHM92 CF3 MezNuuh'll‘NS/Q Me N’| -" 01:3 N \ / CF3 F3C Under an atmosphere of dry N2, 3-prr3’5'CF3 (0.283 g, 1.01 mmol) was loaded into a 20 mL scintillation vial in OEt2 (3 mL). In a separate vial was loaded Ti(NMe2)4 (3) (0.113 g, 0.506 mmol) in OEt2 (3 mL). The two vials were capped and put into a cold well, 3 5—CF3 where they sat until frozen. 3-prrr ’ was added to the vial containing 3 and the solution was allowed to warm to room temperature where it was left to stir for 6 h. After 6 h the solution was concentrated in vacuo to yield a red oil. The oil was then triturated with pentane to yield a red solid. The solid was crystallized from pentane to yield Ti(NMe2)2(NHMe2)(pyrr3’5.C173 )2 as a red solid (0.285 g). Due to probable isomerization of 19 in solution, the 1II and '3C spectra recorded were not consistent with the solid structure. See Appendix for 1H and '3 C spectra for 19. 77 Synthesis of 2-(bis(trifluoromethyl)phenyl)-4-pentafluorophenyl-pyrrole Under an atmosphere of dry N2, [Ir(OMe)(Cod)]2 (0.1g, 0.151 nnnol) was loaded into a 20 mL scintillation vial with hexane (4 mL) and HBpin (1.92 g, 15 mmol) and stirred for 10 min prior to the addition of Butbipy (0.081 g, 0.302 mmol) which was then stirred for another 10 min. To the same vial was added N-Boc-Z-(3,5- (bistrifluoromethyl)phenyl)pyrrole (3.8 g, 10 mmol) in hexane (5 mL). The resulting reaction mixture was then transferred to a Schlenk tube, sealed, taken out of the dry box, and put in a 60 °C oil bath for 36 h. After the reaction was complete by GC-FID, the crude mixture was run through two plugs of silica gel with copious amounts methylene chloride. The volatiles were removed in vacuo to yield a clear oil. Trituration with pentane yielded and N-Boc-2-(3,5-(bistrifluoromethyl)phenyl)-4-pinacolboryl-pyrrole compound as a white solid. The product was used without any further purification. Under an atmosphere of dry N2, a 20 mL scintillation vial was loaded with Pd(PPh3)4 (0.140 g, 0.120 mmol) in DME (5 mL), BrC6F5 (0.632 g., 2.57 mmol), and N-Boc-2-(3,5- (bistrifluoromethyl)phenyl)-4-pinacolboryl-pyrrole (1.18 g., 2.33 mmol). The solution was transferred to a Schlenk tube, sealed, removed from the dry box, and connected to a Schlenk line. Under a continuous flow of N2, the cap was removed, K3POa was added to the vessel, and the vessel was quickly capped. The headspace in the Schlenk tube was 78 evacuated and the tube was heated in an oil bath at 100 0C for 8 d. After 8 d the reaction mixture was transferred to a separatory funnel. The tube was rinsed with H20 (30 mL) and EtOAc (30 mL). The solution was extracted with EtOAc (3 x 30 mL). The combined organic layers were collected and dried over MgSOs. The volatiles were removed in vacuo to yield a viscuous oil. The oil was transferred to a Schlenk tube and placed in a 165 °C oil bath for 10 min under a continuous flow of N2. After 10 min the crude mixture was run through a plug of silica with methylene chloride (500 mL). The volatiles were removed in vacuo to yield a pale pink solid. Crystallization from pentane yielded the product as a pale pink solid (0.2, 19.2 %). M.p 146-149 °C. 1H NMR (500 MHz, CDCl;): 9.01 (br s, 1 H), 7.90 (s, 2 H), 7.72 (s, 1 H), 7.42 (s, 1 H), 7.06 (s, 1 H). ”C {‘H} NMR (125 MHz, CDC13): 133.75, 132.52 (q, JCF = 33.9 Hz), 129.79, 123.18 (q, IQ}? = 276 Hz), 123.8-123.58 (m), 122.12, 122.14-121.9 (m), 120.19-119.5 (m), 111.88, 110.9- 110.05 (m), 109.1-108.85 (m). Anal. (Found) Calcd: C, (48.11) 48.56; H, (1.27) 1.36; N, (3.22) 3.15. 79 Synthesis of Ti(NMez)3(pyr2-CF3-4-C6F5 ) (20) Under an atmosphere of dry N2, Ti(NMe2)4 (3) (0.05 g, 0.223 mmol) was loaded into a 20 mL scintillation vial in OEt2 (2 mL). In a separate vial was loaded 2-(3,5- (bistrifluoromethyl)phenyl)-4-pentafluoropheny pyrrole (0.1 g, 0.223 mmol) in OEt2 (2 mL). The vials were placed in a cold well where that sat until frozen. While still cold, prrrZ'CF3'4'C6F5 was added to the vial containing 3. The solution was allowed to stir at room temperature for ~18 h. The following day the volatiles were removed in vacuo to yield a yellow solid. Crystallization from OEt2/pentane yielded the title compound as a yellow solid (0.12 g 86%). M.p 76-79 °C. 1H NMR (500 MHz, CDC13) 7.85 (2 H, S), 7.70 (1 H, s), 7.37 (1 H, s), 6.80 (1 H, s), 3.13 (18 H, s). ”C {‘H} NMR (125 MHz, CDC13): l45.1-l44.80 (m), 142.98-142.5 (m), 138.45, 139.05-138.73 (m), 137.51- 136.97 (in), 131.46 (q, Jcp = 33 Hz), 131.47-131.25 (m), 126.88 (q, Jcp = 3.5 Hz), 126.75, 123.55 (q, Jcp = 273 Hz), 119.41-119.21 (m), 111.60-111.49 (m), 110.82-110.67 (m), 43.74. Anal. (Found) Calcd: C, (46.25) 46.17; H, (3.73) 3.71; N, (8.86) 8.97. 80 3.5 References 1. 9. (a) Shi, Y.; Hall, C.; Ciszewski, J. T.; Cao, C.; Odom, A. L. Chem. Commun. 2003, 5, 586. (b) Cao, C.; Li, Y.; Shi, Y.; Odom, A.L. Chem. Commun, 2004, 17, 2002. For some recent reviews on hydroamination see: (a) Odom, A. L. Dalton Trans. 2005, 225. (b) Muller, T. E.; Hultzsch, K. C.; Yus, M.; Foubelo, F.; Tada, M. Chem Rev., 2008, 108, 3795. Cao, C.; Shi, Y.; Odom, A. L. J. Am. Chem. Soc. 2003, 125, 2880. . Banerjee, S.; Barnea, E.; Odom, A. L. Organometallics, 2008, 27, 1005. Swartz II, D. L.; Odom. A. L. Organometallics, 2006, 25, 6125. . For some additional reports on iridium catalyzed borylation/coupling reactions see. (a) Paul, S.; Chotana, G. A.; Holmes, D.; Reichle, R. C.; Maleczka, R. E.; Smith III, M.R. J. Am. Chem. Soc. 2006, 128, 15552. (b) Holmes, D.; Chotana, G. A.; Maleczka, R. E.; Smith III, M. R. Org. Lett. 2006, 8, 1407. (c) Chotana, G. A.; Rak, M. A.; Smith III, M. R. J. Am. Chem. Soc. 2005, 127, 10539. (d) Boebel, T.A.; Hartwig, J. F . Organometallics, 2008, 27, 6013. (e) Ishiyama, T.; Takagi, J.; Ishida, K.; Miyaura, N.; Anastasi, N. R.; Hartwig, J. F. J. Am. Chem. Soc., 2002, 124, 390. Heutling, A.; Pohlki, F.; Doye, S. Chem. Eur. J. 2004, 10, 3059. Bradley, D. C.; Thomas, I. M. J. Chem. Soc. 1960, 3859. Espenson, J. H. Chemical Kinetics and Reaction Mechanisms; McGraw-Hill: New York, 1995. Rieth, R. D.; Mankad, N. P.; Calimano, E.; Sadighi, J. P. Org. Lett. 2004, 6, 3981. 10. Balboni, D.; Carnurati, 1.; Prini, G.; Resconi, L. Inorg. Chem. 2001, 40, 6588. 81 CHAPTER 4 Synthesis, structure, and hydroamination kinetics of 3,3’-diaryldipyrrolylmethane titanium complexes 4.1 Introduction Dipyrrolylmethane ligands on titanium are effective catalysts for the hydroamination of primary amines and alkynes as well as iminoamination.l The fastest hydroamination catalyst in the literature as of 2008 was Ti(NMe2)2(dpm) (9), where dpm is 5,5-dimethyldipyrroly1methane (Figure 4.1).2 The previous chapters have discussed the approaches taken to alter the dpm framework to improve catalyst reactivity with hopes of generating more reactive catalysts and applying them to multi-component coupling reactions. However, every attempt to prepare a catalyst more active than 9 has been unsuccessfiil. N N/Ti(NMe2)2 / Ti(NMe2)2(dpm) (9) Figure 4.1 Solid-state structure of Ti(NMe2)2(dpm) (9). Putting sterics on the 2-position of the dpm framework results in decreased hydroamination reactivity, which is attributed to steric inhibition near the substrate binding site.2 In the same report, it was shown experimentally with bis(pyrrolyl) titanium catalysts, that the hydroamination reactivity increases with electron-withdrawing substituents. Moving sterics from the 2-position to the 3—position on the dpm framework could result in faster catalysis than Ti(NMe2)2(dpm) (9). This chapter discusses the synthesis, structure, and hydroamination kinetics of 3,3’-diaryldipyrrolylmethane titanium complexes. 82 4.2 Results and discussions The synthesis of 3-arylpyrrole complexes is greatly enabled by contributions from Smith and co-workers who demonstrated the selective generation of 3-(pinacolboryl)-N- Boc-pyrrole, which can then be coupled with aryl halides using a palladium catalyst. (Scheme 4.1).3 Scheme 4.1 Synthesis of 3-aryl pyrroles. Boc Boc Pd lPhos hine U N l 1 p Boc [Ir] | / Ar—X, Base N + t = Q BPin H Ar 0 \ ’m HBPin Using this synthetic procedure I prepared 3-(3,5-(CF3)-C6H3) PYITOIC (3-HPYIT3’5— CF3) and 3-(2,4,6-(F)3-C6H2) pyrroles (3-prrrC6F3 ) to use as starting materials for 3- substituted dpm derivatives. The decision to use these two pyrroles was driven by a few initial results. First, it was shown experimentally that electron-withdrawing substituents on pyrrolyl ligands increase hydroamination catalysis rates. Therefore, the ligands'should 3,5—CF3 be electron deficient. Second, a comparison of Ti(NMe2)2(pyrr ) (14), which has two CF3 groups in the meta positions of the arene, and Ti(NMe2)2(pyrr4'CF3)2 (16), which has one CF 3 group in the para position on the arene, had similar catalysis rates (Table 4.1). For that reason, I wanted to see what effect placing electron-withdrawing groups at the ortho and para positions had on the catalysis activity. 83 Table 4.1 Comparison of hydroamination catalysis rate constants 3,5—CF 3 Ti(NMe2)2(pyrr )2 (14) and Ti(NMe2)2(pyrr4'CF3)2 (16) from Chapter 2. The reaction conditions are shown in Scheme 4.4. Rates are at the 99% confidence level with at least 3 repeated runs. Entry Catalyst kobs (x104 S-l) F,c CF, 1 1275 :t 72 (Me,N),Ti N \ f — )2 14 CF, 2 1255 :t 145 (MepN)2Tl~CN \ > — 2 16 The usual procedure for the synthesis of 2,2’-diaryldipyrrolylmethanes involves the neat reaction of 2-aryl pyrroles in excess acetone with a catalytic amount of trifluoroacetic acid (TFA). However, using these reaction conditions with 3-aryl pyrroles did not yield the corresponding 3,3’-diary1dipyrrolylmethane derivatives. Switching to a Lewis acid, InC13, catalyzes the reaction of 3-ary1 pyrroles and acetone to yield two new 3,3’-diaryldipyrroly1methanes, 3-H2dpm3’5_CF3 and 3-H2dpmc‘5F3 in 76% and 14% isolated yields. These complexes react with Ti(NMe2)4 (3) to yield two new titanium 84 3,5—CF3 C6133) precatalysts, Ti(NMe2)2(NHMe2)(dpm ) (22) and Ti(NMe2)2(NHMe2)(dpm (23) in 92% and 86% crystallized yields (Scheme 4.2). Scheme 4.2 Synthesis of 3-H2dpm3’5—CF3, 3-H2dme6F3, . 3,5—CF3 . C6F3 T1(NMe2)2(NHMe2)(dpm ) (22) and T1(NMe2)2(NHMe2)(dpm ) (23). CFs CF, ll 0 40/ l Cl 0 °° n 3 F C + excess 4, s ' / A 40 °C, 18 h CF, 74% CF, H F30 3-H2dpm3'5‘CF3 3-prrr3-5'CF3 Ti(NMez)4 (3) 92% 0512 N O 10/ l Cl 00 n 3 + excess .. ' / 2K 40 °C, 18 h F 43% F F Ci-prrrc’fiF3 Ti(NM92)4 (3) 860/0 OEt2 12 h Ti(NM82)2(NHM62)(dme6F3) (23) 85 The solid-state structure of 22 has a pseudo trigonal bipyramidal (tbp) geometry with the pyrrolyls 111,1]1-bound. The 1H NMR spectrum is consistent with the solid-state structure showing equivalent pyrrolyls. There is a broad Singlet at 2.61 ppm in the 1H spectrum that integrates to 6 hydrogens, which are assigned to the methyl groups on dimethylamine. As expected, the longest Ti-N bond length is for the dimethylamine ligand at 2.135(17) A, which sits in the equatorial plane. Attempts to grow X-ray quality crystals of 23 are currently underway. The 'H NMR Spectrum of 23 is similar to 22, showing equivalent pyrrolyls and a broad singlet at 2.49 ppm corresponding to 6 hydrogens, which are assigned to the methyl groups on a coordinated dimethylamine. 86 H532) F(531) F(551),,. ' F(553) . 8' 3 ~‘ "“‘F(533) ‘13 a F(552) 1' ,, Q Q Q 0 1:“(433) . . 321,045,) “185131 ‘5 1' ,6, ( )F(451’ Figure 4.2 Solid-state structure of Ti(NMe2)2(NHMe2)(3-dpm3’5_CF3) (22) from single crystal X-ray diffraction. Selected bond distances (A) and bond angles (deg): Ti- N(1) 1.931(18), Ti-N(2) 1.826(13), Ti-N(3) 2.135(17), Ti-N(4) 2.051(9), Ti-N(5) 2.067(9), N(1)-Ti-N(2) 109.6(9) N(2)-Ti-N(3) 97.7(7), N(3)-Ti-N(4) 84.8(5), N(4)-Ti- N(5) 82.3(4), N(2)-Ti-N(4) 108.8(6), N(2)-Ti-N(5) 102.8(5), N(3)-Ti-N(5) 158.4(6), N(1)-Ti-N(3) 87.2(7), N(1)-Ti-N(4) 141.4(8), N(1)-Ti-N(5) 92.2(6). To test the kinetic ability of these catalysts compared to Ti(NMe2)2(dpm) (9), precatalysts 22 and 23 were tested under the kinetic conditions reported in Chapter 2 (Scheme 4.3).2 In addition to running the reactions at 75 °C in toluene-d3 in the NMR probe, they were also run at 75 °C in a reaction calorimeter.4’5 The disappearance of the 87 l-phenylpropyne starting material versus time was used to fit the first-order equations as measured by 1H NMR spectroscopy. Alternatively, integrated heat flow versus time, t, divided by the total heat of reaction would be used to give the percent conversion, which also fit well to first order equations.6 The methods were consistent with each other, producing the same rate for each catalyst. The results of the kinetic studies on compounds 9, 22, and 23 are shown in Table 4.2. The errors are based of at least three repeated runs and are at the 99% confidence level. The errors varied from as little as 7% to as much as 10%. Scheme 4.3 Reaction conditions for kinetic study Me 10% catalyst 0.05 M NPh 10 NH2Ph + é 4, Ph/ toluene-d8, 75 °C /u\/ Ph 5 M 0.5 M d 1- hen ro ne 1 P th'P DY l = kobst From the data in Table 4.2, 3-aryl substitution on the dpm results in an increase in rate constant by a factor of about 3.5. As anticipated, adding electron-withdrawing groups“ without increasing steric congestion around the metal results in catalysis faster than Ti(NMe2)2(dpm) (9). The reaction of aniline and l-phenylpropyne is smoothly catalyzed by 22 and 23 in ~1 h with rate constants of (6963 s 582) and (6225 a 614) x 10‘7 s“. 88 Table 4.2. Representative catalysis rates for Ti(NMe2)2(dpm) (9), Ti(NMe2)2(NHMe2)(3-dpm3’5“CF3) (22) and Ti(NMe2)2(NHMe2)(3-dme6F3) (23). Entry cataIYSt kObS (X104 S-l)a 1 Ti(NMe2)2(dpm) 1976 :l: 130 9 2 Ti(NMe2)2(NHMe2)(3-dpm3’5_CF3) 6963 :1: 582 22 3 Ti(NMe2)2(NHMe2)(3-dme6F3) 6225 t 614 23 a Errors are based at 99% confidence level with at least 3 repeated runs. 89 4.3 Conclusions Using readily prepared 3-substituted pyrroles, a route to 3,3’- diaryldipyrrolylmethanes has been developed where the synthesized pyrroles can be used as the limiting reagent in condensation with acetone. Placing these new ligands on titanium is readily accomplished by reaction with Ti(NMe2)4 (3), and the resulting complexes Show nl,nl-coordination of the dipyrrolylmethane in the solid-state. The increased electrophilic nature of the metal center allows for dimethylamine retention. Catalysis with these new complexes was faster than the dpm parent complex 9, suggesting that electron-withdrawing groups increase hydroamination reactivity. However, strategic positioning of the electron-withdrawing groups did not result in a significant difference in catalysis rate, which can be attributed to the choice of electron- withdravving substiuents. While CF3 and F are good o-withdrawing groups, it is possible for F to be a rt-donor negating some of the electron-withdrawing ability. Investigations are currently underway with using just CF 3 substitution in ortho/para positions on the arene to see if there is a Significant difference in catalysis rate. 90 4.4 Experimental General Considerations: All manipulations of air sensitive compounds were carried out in an MBraun drybox under a purified nitrogen atmosphere. Anhydrous ether was purchased from Columbus Chemical Industries Inc. Pentane and toluene were purchased from Spectrum Chemical Mfg. Corp., were purified by sparging with dry N2, then dried by running through activated alumina systems purchased from Solv-Tek. Hexanes and ethyl acetate were purchased from Mallinckodt Baker Inc. Pinacolborane was purchased from BASF and was used as received. 1-bromo-3,5- bis(trifluoromethyl)benzene and 2-bromo-l,3,5-trifluorobenzene were purchased from Matrix Scientific and dried by passed over a column of neutral alumina. N-Boc-pyrrole was purchased from Aldrich, vacuum distilled, and stored under a purified nitrogen atmosphere over molecular sieves. Aniline was purchased from Matheson Coleman and Bell Mfg. and was distilled twice from calcium hydride under vacuum. l-phenylpropyne was purchased from GFS Chemical, distilled under vacuum, and then passed over two columns neutral alumina. Ti(NMe2)4,7 was synthesized according to the literature procedure. Deuterated solvents were dried over purple sodium benzophenone ketyl (C5D6) or phosphoric anhydride (CDC13) then distilled under a nitrogen atmosphere. Deuterated toluene was dried by passing it through two columns of neutral alumina. 1H and 13 C spectra were recorded on Inova—300 or VXR-500 spectrometers. All spectra were referenced internally to residual protiosolvent ('H) or solvent (13C) resonances. General Considerations for X-ray Crystallography: Crystals grown from concentrated solutions in pentane at —35 °C were moved quickly from a scintillation vial to a microscope slide containing Paratone N. Samples were selected and mounted on a 91 glass fiber in wax and Paratone. The data collections were carried out at a sample temperature of 173 K on a Bruker AXS platform three-circle goniometer with a CCD detector. The data were processed and reduced utilizing the program SAINTPLUS supplied by Bruker AXS. The structures were solved by direct methods (SHELXTL v5.1, Bruker AXS) in conjunction with standard difference Fourier techniques. General Considerations for Kinetics: All manipulations were done in an inert atmosphere drybox. In a 2 mL volumetric flask was loaded the catalyst (10 mol%, 0.1 mmol), aniline (0.931 g, 911 11L, 10 mmol), l-phenylpropyne (0.116 g, 125 11L, 1 mmol), and ferrocene (0.056 g, 0.3 mmol) as an internal standard. The solution was then diluted to 2 mL with deuterated toluene. An ample amount of solution (~0.75 mL) was put into a threaded J. Young tube that was sealed with a cap and then wrapped with Teflon tape. The tube was then removed from the drybox and heated at 75 °C in the NMR spectrometer. The relative I-phenylpropyne versus ferrocene concentration was monitored as a function of time. For kinetics by reaction calorimetry, the same amounts of reagents were used. The solution was added to the reaction vial and sealed with a Teflon cap minus 1- phenylpropyne. The vial was taken out of the dry box and put in the calorimeter at 75 °C. Once the solution had reached 75 °C, the alkyne was added via syringe. The progress of the reaction was monitored by heat as a function of time, and the kinetic rates were verified by 1H NMR spectroscopy. The fits are to the exponential decay of the starting material or exponential growth of the heat using the scientific graphing programs Origin or KaleidaGraph. The exact _ t expression used to fit the data was Y, = Yo, + (Y0 - Yoo)exp kObs where Y = [1- 92 phenylpropyne] at time = t (Y 1), infinity (1’00), or at the start of the reaction (Y0).8 The variables Yoo, Y0, and kobs were optimized in the fits. 93 Synthesis of 3-(3,5-bistrifluromcthylphenyl-lH-pyrrole) (3-prrr3’5'CF3) \ZI CF3 F30 Under an atmosphere of dry N2 an oven dried 250 mL Schlenk tube was loaded with a stirbar, Pd(PPh3)4 (0.536 g, 0.464 mmol) in DME (10 mL), 3,5- bis(trifluoromethyl)phenylbromobenzene (4.54 g, 15.5 mmol) and N-Boc-3- pinacolborylpyrrole (4.54 g, 15.5 mmol) in DME (20 mL). To that same vessel was added K3PO4 (6.16 g, 23.15 mmol). DME (7 mL) was used to wash the sides of the tube. The tube was then capped, removed from the drybox, and placed in an oil bath at 85 °C for 18 h. After the reaction was complete as judged by GC-FID, the reaction mixture was transferred to a separatory funnel with H20 (100 mL). The aqueous layer was extracted and washed with ethyl acetate (3 x 75 mL). The combined organic layers were combined and dried with MgSO4_ The volatiles were removed in vacuo to yield a viscous brown oil. The oil was then transferred to that same 250 mL Schlenk tube with l-butanol (35 mL) and K3PO4 (6 g, 23 mmol). The solution was then heated in an oil bath at 100 °C for 18 h. The resulting solution was then transferred to a separatory funnel with H20 (75 mL) water. The aqueous layer was extracted and washed with ethyl acetate (3 x 75 mL). The combined organic layers were collected and dried with MgSOa. The volatiles were removed in vacuo to yield a viscous brown oil. The product was then purified by column chromatography on basic alumina with an elution of hexaneszethyl acetate (4:1) to yield the product as an off white solid. The product was purified further by sublimation to give 94 a bright white solid (2.7 g, 62.5 %). lH NMR(CDC13, 500 MHz): 8.37 (br s, l H), 7.90 (s, 2 H), 7.65 (s, 1 H), 7.18 (m, 1 H), 6.87 (q, l H), 6.58 (m, 1 H). ”C {‘H} NMR (CDC13, 125 MHz): 138.04, 131.81 (q, 1,, = 31.51 Hz), 124.85 (dd), 123.77 (q, JCF = 277.72 Hz), 122.53, 119.78, 118.77 (sep., 1c, = 4.63 Hz), 115.74, 106.63. Anal. (Found) Calcd: C, (51.63) 51.29; H, (2.53) 2.33; N, (5.02) 5.03. 95 C8F6 Synthesis of 3-(2,4-bis(trifluoromethyl)phenyl)-lH-pyrrole (prr ) \ZI CF, CF3 Under an atmosphere of dry N2, a 200 mL Schlenk tube was loaded with 3- (pinacolboryl)-N-Boc-pyrrole (4.86 g, 16.5 mmol), K3P04 (anhydrous) (3.51 g, 16.5 mmol), Pd2(dba)3 (0.216 g, 0.236 mmol) biphenyl-2-yldicyclohexylphosphine (0.165 g, 0.471 mmol) and a stirbar. The Schlenk tube was capped, removed from the drybox, and connected to a Schlenk line. The headspace was evacuated and back filled with N2. The cap was removed quickly under a continuous flow of N2 and replaced with a septum. A 60 mL syringe containing t-amyl alcohol (30 mL) and 1-bromo-2,4- bis(trifluoromethyl)benzene (3.46 g, 11.8 mmol) was inserted into the septum with the needle extending below the neck of the Schlenk tube, and the solution was added. The septum and needle were removed quickly, and the tube was sealed with a screwcap. The headspace was evacuated and filled with N2. This procedure was repeated 6 times. The Schlenk tube was then placed in a 100 °C oil bath for 18 h. Once the reaction was complete as judged by GC-FID, K3P04 (3 g, 14 mmol) and Bu"OH (OH) were added. The reaction mixture was allowed to reflux for an additional 2 h at 100 °C. After the deprotection, the resulting solution was poured into a separatory funnel with H20 (100 mL) and ethyl acetate (50 mL). The water layer was washed and extracted with ethyl acetate (3 x 75 mL). The combined organic layers were collected and dried over MgSOa. The volatiles were removed in vacuo to give a viscous red oil. The product was then 96 purified by column chromatography on silica gel using an elutant of CH2Cl2zhexanes (7:3). This was followed by a second column on silica gel with hexaneszethyl acetate (7:3) as elutants to give the product as a pale orange solid (2.6 g, 76%). M.p 27 - 30 °C 1H NMR (CDC13, 500 MHz): 8.35 (br s, l H), 7.99 (s, 1 H), 7.75 (d, 1 H, Jim = 8.14 Hz), 7.62 (d, 1 H, JHH = 8.14 Hz), 7.00 (s, 1H), 6.87-6.85 (m, 1 H), 6.45 (app s, 1 H). ”C {‘H} NMR (CDC13, 125 MHz): 140, 132.79, 128.28 (q, Jcp = 33 Hz), 128.34 (q, Jcp = 30 Hz), 128.14-128.05 (m), 123.87 (q, Jcp = 278 Hz), 123.77 (q, Jcp = 271 Hz), 123.66-123.49 (m), 120.93, 118.30, ll7.91-117.83 (m), 10988-10983 (m). Anal. (Found) Calcd: C, (51.26) 51.63; H, (2.42) 2.53; N, (4.87) 5.02. 97 Synthesis of 3-(2,4,6-trifluorophenyl)—lH-pyrrole (3-prrrC6F3) Under an atmosphere of dry N2 a 200 mL Schlenk tube was loaded with 3-(pinacolboryl)- N-Boc-pyrrole (4.16 g, 14.2 mmol), K3PO4 (anhydrous) (2.81 g, 13.2 mmol), Pd2(dba)3 (0.156 g, 0.173 mmol), biphenyl-2-yldicyclohexylphosphine (0.119 g, 0.346 mmol) and a stirbar. The Schlenk tube was capped, removed from the drybox, and connected to a Schlenk line. The headspace was evacuated and back filled with N2. The cap was removed quickly under a continuous flow of N2 and replaced with a septum. A 60 mL syringe containing t-amyl alcohol (30 mL) and 2-bromo-1,3,5-trifluorobenzene (2 g, 9.5 mmol) was inserted into the septum with the needle extending below the neck of the Schlenk tube, and the solution was added. The septum and needle were removed quickly, and the tube was sealed with a screwcap. The headspace was evacuated and filled with N2. This procedure was repeated 6 times. The Schlenk tube was then placed in a 100 °C oil bath for 18 h. Once the reaction was complete as judged by GC-FID, K3P04 (2 g, 9.4 mmol) and Bu"OH (8 mL) were added. The reaction mixture was refluxed for an additional 2 h at 100 °C. After the deprotection, the resulting solution was poured into a separatory firnnel with H20 (50 mL) and ethyl acetate (50 mL). The water layer was washed and extracted with ethyl acetate (3 x 75 mL). The combined organic layers were collected and dried over MgSOa. The volatiles were removed in vacuo to give a viscous red oil. The flask containing the oil was connected to the Schlenk line and kept under 98 vacuum for 24 h. The product was then purified by column chromatography on silica gel using of CH2Cl2zhexanes (7:3) as elutants. A second column on silica gel with of hexaneszethyl acetate (4:1) as elutants gave the product as an off white solid, (0.958 g, 52%). 1H NMR (CDC13, 500 MHz): 8.38 (br s, 1 H), 7.28-7.27 (m, 1 H), 6.90-6.78 (m, 1 H), 6.77-6.71 (m, 3 H). 13C {H} NMR (CDC13, 125 MHz): 161.14-160.65 (m), 159.16- 158.69 (m), 118.74 (app t), 117.89, 110.76, 109.40 (app t), 100.54-100.09 (m). Anal. (Found) Calcd: C, (60.50) 60.92; H, (3.14) 3.07; N, (6.84) 7.10. 99 Synthesis of 5,5'{propane-2,2-diyl)bis(3-(3,5-bis(trifluoromethyl)phenyl)-1H- pyrrole) (3-H2dpm3’5'CF3) An oven dried pressure tube was loaded with 3-(3,5-bistrifluromethylphenyl-lH-pyrrole) (2.33 g, 8.35 mmol), acetone (13 mL, 10.27 g, 176 mmol), and a stirbar. To that tube was added InC13 (0.3 g, 1.3 mmol). The tube was then capped and placed in an oil bath at 40 °C for 24 h. Once the reaction was complete as judged by 1H NMR spectroscopy, the solution was transferred to a 250 mL round bottom flask, and the volatiles were removed in vacuo. The resulting solid was put into a solution of hexaneszCH2Cl2 (7:3) (~35 mL). A column was made using hexanes:CH2Cl2 (7:3). Once the crude mixture was loaded onto the column, it was flashed with hexanes until spots appeared on the TLC plate. Once material started appearing on TLC the elutant was changed to hexaneszCH2Cl2 (7:3). Once the desired product appeared by TLC, the elutant was changed to just CH2Cl2 until all the product had been collected. The fractions were collected, and the volatiles were removed in vacuo to yield the product as a white solid, (1.9 g, 76 %) lH NMR (CDC13, 500 MHz): 8.03 (br S, 2 H), 7.87 (s, 4 H), 7.62 (s, 2 H), 7.06 (q, 2 H), 6.45 (q, 2 H), 1.74 (s, 6 H). ”C {'H} NMR (CDC13, 125 MHz): 140.67, 138.13, 132.09 (q, J,F = 32.14 Hz), 124.88, 123.77 (q, Jcp = 272.73 Hz), 122.46, 118.99 (pent, JCF = 4.11 Hz), 115.61, 102.73, 35.85, 29.19. Anal. (Found) Calcd: C, (53.88) 54.14; H, (3.31) 3.03; N, (4.51) 4.68. 100 Synthesis of 5,5'-(propane—Z,Z-diyl)bis(3-(2,4,6-trifluorophenyl)-lH—pyrrole) (3- szme6F3) A threaded pressure tube was charged with InCl3 (0.107 g, 0.0484 mmol), 3-(2,4,6- trifluorophenyl)—lH-pyrrole (0.952 g, 0.483 mrnol), acetone (2.80 g, 3.55 mL, 4.83 mmol), and a stirbar. The tube was capped and placed in a 40 °C oil for 18 h. Once the reaction was complete as judged by 1H NMR spectroscopy, the solution was transferred to a 100 mL round bottom flask, where the volatiles were removed in vacuo to yield a viscous orange oil. The product was purified by column chromatography on silica gel using hexaneszethyl acetate (85:15) as elutants. The product was obtained as an off white solid, (0.450 g, 43 %). 1H NMR (500 MHz, CDC13): 7.09 (s, 2H), 7.07-7.06 (m, 2 H), 6.75-6.65 (m, 4 H), 6.55-6.53 (m, 2 H). 13C {H} NMR (125 MHz, CDC13): 161.12- 160.67 (m), 159.12-158.72 (m), 138.68, 118.37 (1), 110.42, 110.05-109.83 (111), 105.12- 105.01 (m), 100.63-100.15 (m), 35.35, 28.96. Anal. (Found) Calcd: C, (63.11) 63.60; H, (3.58) 3.31; N, (6.17) 6.48 101 Synthesis of Ti(NMe2)2(NHMe2)(3-dpm3’5‘C”) (22) 3,5—CF3 Under an atmosphere of purified N2, a vial was loaded with 3-H2dpm (0.503 g, 0.841 mmol) in OEt2 (3 mL). A separate vial was loaded with Ti(NMe2)4 (3) (0.188 g, 0.840 mmol) in OEt2 (3 mL). The two vials were placed in the cold well where they sat 3,5-CF3 until frozen. To the thawing solution, 3-H2dpm was added to the vial containing 3. The solution was allowed to stir overnight. The next day, the volatiles were removed under reduced pressure to give a red oil. The oil was triturated with pentane thrice to give the product as a bright orange solid. The solid was crystallized from OEt2/pentane to yield the title compound as an orange solid, (0.602 , 92%). 1H NMR (CDC13, 500 MHz): 7.83 (s, 4 H), 7.53 (s, 2 H), 7.07 (s, 2 H), 6.40 (s, 2 H), 3.37 (s, 12 H), 2.61 (br s, 6 H), 1.72 (s, 6 H). 13C {'H} NMR (CDC13, 125 MHz): 154.05 (m), 138.8, 131.5 (q,Jcp = 34 Hz), 124.4, 123.8 (q, Jcp = 270 Hz), 122.2, 121.8, 117.6, 100.9, 46.8, 40.5, 38.2, 31.1. Anal. (Found) Calcd: C, (50.75) 50.98; H, (4.89) 4.54; N, (9.04) 9.01. 102 Synthesis of Ti(NMe2)2(NHMe2)(3_dme6F3 ) (23) Under an atmosphere of purified N2, a vial was loaded with 3-H2dpmc6F3 (0.130 g, 0.3 mmol) in OEt2 (3 mL). A separate vial was loaded with Ti(NMe2)4 (3) (0.067 g, 0. mmol) in OEt2 (3 mL). The two vials were placed in the cold well, where they sat until frozen. To the thawing solution, 3-H2dmeéF3 was added to the vial containing 3. The solution was allowed to stir overnight. The next day the volatiles were removed under reduced pressure to give a red oil. The oil was triturated with pentane twice to give the product as a bright orange solid. The solid was crystallized from OEt2/pentane to yield the title compound as an orange solid, (0.157 g, 86%). M.p 110-112 °C. 1H NMR (CDCl3, 500 MHz): 7.32 (s, 2 H), 6.72-6.64 (m, 6 H), 6.63 (s, 2 H), 3.24 (s, 12 H), 2.49 (br s, 6 H), 1.73 (s, 6 H). 103 4.5 References 1. (a) Shi, Y.; Hall, C.; Ciszewski, J. T.; Cao, C.; Odom, A. L. Chem. Commun. 2003, 586. (b) Ramanathan, B.; Keith, A.J.; D. Armstrong D.; Odom, A.L. Org. Lett., 2004, 6, 2957. (c) Majumder, S.: Odom, A.L.; manuscripts in preparation. 2. Swartz, D.L.; Odom, A.L. Organometallics, 2006, 26, 6125. 3. (a) Tse M.K.; Cho, J-Y.; and Smith, 111 MR. Org. Lett., 2001, 3, 2831. (b) Paul, S.; Chotana, G.A.; Holmes, D.; Reichle, R.C.; Maleczka, R.E.; Smith 111, MR. J. Am. Chem. Soc. 2006, 128, 15552. . For a nice review on kinetics obtained using reaction calorimetry see. Blackmond, D. G. Angew. Chem, Int. Ed. 2005, 44, 4302. . For some additional reports on reaction calorimetry used in catalysis see. (a) Antonio Jinu C. F.; Mathew, J. S.; Blackmond, D. G.; Ind. Eng. Chem. Res, 2007, 46, 8584. (b) Mathew, J. S.; Klussmann, M.; Iwamura, H.; Valera, F.; Futran, A.; Emanuelsson, E.A.C.; Blackmond, D. G.; J. Org. Chem, 2006, 71, 4711. (c) Blackmond, D. G.; Ropic, M.; Stefinovic, M. Org. Process Res. Dev., 2006, 10, 457. (d) Shekhar, S.; Ryberg, P.; Hartwig, J. F.; Mathew, J. S.; Blackmond, D. G.; Strieter, E. R.; Buchwald, S. L. J. Am. Chem. Soc., 2006, 128, 3584. 6. Blackmond, D.G.; Rosner, T.; Pfaltz, A. Org. Process Res. Dev., 1999, 3, 275. 7. Bradley, D.C.; Thomas, I.M.; J. Chem. Soc., 1960, 3859. 8. Espenson, J. H. Chemical Kinetics and Reaction Mechanisms; McGraw-Hill: New York, 1995. 104 CHAPTER 5 Effects of 5,5-substitution on dipyrrolylmethane ligand isomerization 5.1 Introduction Dipyrrolylmethane ligands on titanium provide catalysts capable of extremely fast hydroamination of alkynes with primary amines 2and very efficient catalysts for iminoamination.3 One of the most commonly employed dipyrrolylmethane derivatives is prepared from acetone and pyrrole in the presence of trifluoroacetic acid.4 The product 5,5-dimethyldipyrrolylmethane (H2dpm) can be placed on titanium by transamination with Ti(NMe2)4 (3) to provide Ti(NMe2)2(dpm) (9) (Scheme 5.1). 1’2 Scheme 5.1 Synthesis of H2dpm and Ti(NMe2)2(dpm) (9). H o 10/ TFA ”N \ Ti(NMe) (a) on 24 2° = = T' NM U + )K 5min°,/ RT [H 01:12 / N/ '( e,), N/ 90% , H2dpm T1(NMe2)2(dpm) (9) In the solid state and in solution, 9 has a structure where the two pyrrolyl substituents are inequivalent. One of the pyrrolyl rings adopts an 115-geometry, and the other is til-bound. However, at room temperature in the 1H NMR spectrum the two pyrroles appear equivalent due to fast exchange on the timescale of this spectroscopy. Our research group previously reported the barrier for pyrrolyl exchange as AGI ~ 10 heal/mol.l 105 1,3 diaxial interactions Figure 5.1 Use of 1,3-diaxial interactions to affect dipyrrolylmethane isomerizations barriers. The most likely mechanism for pyrrole exchange is conversion of both rings to an Til-geometry} Methods for altering this barrier may provide clues for the active species in catalyses and allow control of complex structure. Previously,6 we have reported that altering the pyrrolyl ligand by placing aryl- substituents in the 2-position of the dpm framework lowers the barriers associated with pyrrolyl exchange, presumably by sterically destabilizing the 115-bonding mode.7 In fact placing a 3,5-bis(trifluoromethyl)phenyl or mesityl group in the 2-position of the pyrrole lowers the pyrrolyl exchange barriers below what can be measured by variable temperature NMR (<5 kcal/mol). However, addition of sterically hindered groups in these positions hinders access to the metal center as well, affecting catalysis {31637 106 Scheme 5.2 Synthesis of H2cpm, H2tmcpm, Ti(NMe2)2(cpm) (24) and Ti(NMe2)2(tmcpm) (24). 0 Ft Ll “(>9 10% TFA R R R excess IL) + > H H / R R N N \ I l / H: Hchm, 43% Me: Hztmcpm, 24% Ti(NMez)4 o t, /Ti’\"NM62 NM62 / Z R = H: T1(NM62)2(Cpm) (24) R = Me: Ti(NMez)2(tmcpm) (25) In an attempt to alter the pyrrolyl isomerization rates without blocking access to the metal I explored the use of substituents in the dpm backbone. The strategy was to use 1,3-diaxial interactions within a cyclohexyl ring to stabilize the 115-bonding mode by sterically inhibiting pyrrolyl-ring rotation (Figure 5.1). The steric interaction will be with the axial pyrrolyl substituent, which will prefer to be in the 115-bonding mode allowing it to present one side of the its-system to the cyclohexyl substituents rather than the larger ring edge. By using gem-disubstitution, the 1,3-diaxial interactions are required regardless of which ring is bound through the rr-face. The Size of R should affect the ease with which the 115-bound ring can rotate to obtain the rill-bonding mode. This chapter discusses the synthesis of 1,1-bis(a-pyrrolyl)cyclohexane ligands, their titanium complexes, and the barriers for pyrrolyl exchange. The barriers were 107 determined using the Eyring method with kinetics from variable temperature spin saturation magnetization transfer NMR spectroscopy. Consequently, the enthalpic and entropic effects of this substitution were determined. All this data will be compared with the Simple dimethyl compound 9. 108 5.2 Ligand Isomerization Study I examined two cyclohexyl-dpm derivatives, which were easily prepared using commercially available cyclohexanone and 3,3,5,5-tetramethylcyclohexanone as shown in Scheme 5.2. The 1,1-bis(a-pyrrolyl)cyclohexane (H2cpm) and 1,1-bis(a- pyrrolyl)-3,3,5,5-tetramethylcyclohexane (H2tmcpm) compounds were prepared in 43% and 24% isolated yields. The transamination reactions with Ti(NMe2)4 (3) and both of these cyclohexyl derivatives occurs in high yields to provide R = H Ti(NMe2)2(cpm) (24) and R = Me Ti(NMe2)2(tmcpm) (25). The structure of the tetramethylcyclohexyl complex was determined by X-ray diffraction. An ORTEP diagram is shown in Figure 5.2. The metric parameters around titanium are quite similar to previously reported derivatives."2 The cyclohexyl group has the expected chair-conformation with the iris-pyrrolyl in the more sterically favorable axial position. This conformation puts the two axial methyl groups directly over the ins-pyrrolyl. 109 Figure 5.2 ORTEP diagram of the X-ray diffraction model for Ti(NMe2)2 (trncpm) (25). Ellipsoids at the 50% probability level. Hydrogens are omitted for clarity. Selected bond distances (A) and angles (°): Ti(1)-N(1) 2.292(2), Ti(1)-N(2) 2.023(1), Ti(l)-N(3) 1.901(1), Ti(1)-N(4) 1.895(2); N(4)-Ti(1)-N(3) 103.49(6), N(4)-Ti(1)-N(2) 107.01(6), N(3)-Ti(1)—N(2) 101.47(6). The increased steric interaction between the axial cyclohexane substituent and the 115-pyrrolyl Jr-face raises the barrier for pyrrolyl exchange as can be seen in the 1H spectrum at room temperature. There is significant line broadening in the room temperature lH spectrum, which indicates that the coalesense point is nearly reached at this tempertature. Cooling a solution of 24 in CDCl3 to —40 °C provided a spectrum of the 111,115 -complex. The 1H NMR spectra of 24 at room temperature and at —40 °C is shown in Figure 5.3. 110 Variable temperature spin-saturation magnetization transfer was used to determine the barriers for pyrrolyl equilibration for Ti(NMe2)2(dpm) (9), Ti(NMe2)2(cpm) (24), and Ti(NMe2)2(tmcpm) (25). As an example, the Eyring plot for Ti(NMe2)2(tmcpm) (25) is shown in Figure 5.4. Figure 5.3. 1H spectra of 24 at room temperature and —40 °C in CDC13. 25 °C ‘ I .. 4 1 d J 4 1 q C q u- N-J — —40 °C lll Figure 5.4 Eyring plot of pyrrolyl exchange in Ti(NMe2)2(tmcpm) (25). I —y-m.107-M.OI RIO“ .3’ ......... ..........1.z--. 0.” 0.0030 0.004 0.0041 0.0042 GM 1” The plot in Figure 5.4 is of ln(k/1') versus UT and is used with the Eyring Equation (Equation 5.1), where k3 = Boltzmann constant and R is the gas constant, to determine the activation parameters under the usual assumptions of Transition State Theory.8 Consequently, the slope (m) can be used to determine AHI using Equation 5.2, and the intercept (b) can be used to determine ASI using Equation 5.3. ln(.]_(. AS: + In .k_B _ E T R 11 RT (5'1) Ali!“t a-me (5.2) A51 -R[b-ln(kTB)] (5.3) The parameters for the pyrrolyl exchange in compounds 9, 24, and 25 can be found in Table 5.1. The enthalpic barriers increase from approximately 12 kcal/mol to over 17 kcal/mol for 9, 24, and 25 reflective of the greater steric interactions on attempting to rotate the 115-pyrrolyl substituent during the exchange in the 112 tetramethylcyclohexyl complex. Table 5.1 Parameters for pyrrolyl exchange for 9, 24, and 25 at 25 °C. AH“ As“ AG“ (kcal/mol) (cal/mol0K) (kcal/mol) Ti(NMe2)2(dpm) 1 1.8 5 .2 10.3 (9) Ti(NMe2)2(cpm) 14.1 5.1 12.6 (24) Ti(NMe2)2(tmcpm) 17.2 12.7 13.4 (25) The entropic parameters are the same, small and positive, for the 5,5- dimethyldipyrrolyl 9 and cyclohexyldipyrrolyl 24 during the exchange. The tetramethyl complex 25 shows a substantial increase in AS1 to 13 cal/mOIOK, which is quite a large positive value for a unimolecular process. The increase in entropy in the transition state suggests that the ground state is highly ordered. This may be due to the single preferred conformation when one of the pyrroles is in the ins-conformation as opposed to the more conformationally flexible nl-conformation where neither chair conformer will be strongly favored and any C—C rotations difficult in the 111,115- conforrner will be more readily allowed. AS Shown in Table 5.1, the difference in exchange barriers between 9 and 24 are largely enthalpic and presumably caused by the 1,3-diaxial interactions. In contrast, the difference in barriers between, 9 and 25, is more complex. The enthalpic barrier to reach the transition state for exchange is ~5 kcal/mol higher for 3, but the entropic term reduces the overall barrier to 13.4 kcal/mol. 113 5.3 Conclusion Cyclohexyl substituents in the backbone of the dipyrrolylmethane ligand can be used to affect pyrrolyl exchange barriers. The difference in sterics for the 115- pyrrolyl and til-pyrrolyl groups can be used to alter this barrier. While placing gem- dimethyl groups on the 3-carbon of the cyclohexyl ring increases the enthalpic cost to pyrrolyl exchange substantially (increased ~50% relative to 9), the entropic factor for sterically hindered 25 cancels some of the enthalpic increases. Consequently, the free energy barrier only increased by a modest 3 kcal/mol using this strategy. All three of these compounds were evaluated for catalysis using the same hydroamination test conditions used previously.6 All three exhibited the same rate for catalysis, indicating that altering the isomerization barriers to the level possible here did not affect the catalysis rates, which is consistent with the rate limiting step not being associated with this isomerization. This is not unexpected as barriers associated with the isomerization are relatively small for this titanium system. The same methodology used here may prove useful to study pyrrolyl coordination effects in systems where the isomerization has a much larger enthalpic barrier. In addition, this strategy using cyclohexane-based dipyrrolylmethanes potentially could be used to control complex chirality by using asymmetric cyclohexane substituents in systems with a larger tendency to stay in the 111,115 -coordination mode. 114 5.4 Experimental General Considerations: Anhydrous ether was purchased from Columbus Chemical Industries Inc., and pentane and toluene was purchased from Spectrum Chemical Mfg. Corp., were purified by sparging with dry N2, then water was removed by running through activated alumina systems purchased from Solv-Tek. Hexanes and ethyl acetate were purchased from Mallinckodt-Baker Inc. Reagent grade cyclohexanone and 3,3,5,5-tetramethylcyclohexanone were purchased from Acros Organics and used as received. Pyrrole was purchased from TCI, was refluxed with sodium, distilled under nitrogen, and then was stored in a purified nitrogen glove box. Ti(NMe2)49 was prepared using a modification of the literature procedure. Deuterated solvents were dried over purple sodium benzophenone ketyl (C6D6) or phosphoric anhydride (CDC13) and distilled under a nitrogen atmosphere. 1H and 13 C spectra were recorded on a VXR-SOO spectrometers. All spectra were referenced internally to residual protiosolvent ('H) or solvent (13C) resonances. Chemical shifts are reported in ppm, and coupling constants are reported in Hz. Typical coupling constants are not reported. Procedure for Spin Saturation Transfer Experiments: The spin saturation transfer experiments were carried out using the method described by Kresge and co- workers.10 The dimethylamido resonances were observed in the spin saturation transfer experiments. In order to correct for decoupler spill-over, off-resonance irradiation was carried out on the opposite side of the observation peak relative to the dimethylamido resonance being irradiated. General Considerations for X-Ray Diffiaction: Crystals grown from 115 concentrated toluene solutions at —35 °C were moved quickly from a scintillation vial to a microscope slide containing Paratone N. Samples were selected and mounted on a glass fiber in wax and Paratone. The data collections were carried out at a sample temperature of 173 K on a Bruker AXS platform three-circle goniometer with a CCD detector. The data were processed and reduced utilizing the program SAINTPLUS supplied by Bruker AXS. The structures were solved by direct methods (SHELXTL v5.1 , Bruker AXS) in conjunction with standard difference Fourier techniques. 116 Synthesis of l,1-bis(a.-pyrroly1)cyclohexane (H2cpm) H H \ I \ / An oven dried 100 mL round bottom flask was charged with cyclohexanone (1 g, 10 mmol) and pyrrole (17 g, 253 mmol) and capped with a septum. The solution was then degassed with argon for 10 min. Trifluoroacetic acid (0.116 g, 1 mmol) was added via syringe. The reaction mixture was allowed to stir for 15 min under an argon atmosphere before being quenched with a 0.1 M NaOH (30 mL) solution. The solution was then transferred to a separatory funnel. It was extracted with OEt2 and the aqueous layer was washed with OEt2 (2 X 30 mL). The combined organic layers were dried with MgSOs and subjected to rotary evaporation to yield a viscous brown oil. The excess pyrrole was removed by distillation under vacuum (~l torr). The product was purified by column chromatography on silica gel with an eluant of hexaneszethyl acetate (7:3) to yield a white solid, (0.928 g, 43%). M. p. 104-106 °C. 1H NMR (CDC13, 500 MHz): 7.61 (br s, 2 H), 6.58-6.55 (m, 2 H), 6.18-6.13 (m, 2 H), 6.13-6.11 (m, 2 H), 2.12-2.07 (m, 4 H), 1.62-1.53 (m, 4 H), 1.52-1.45 (m, 2 H). 13C {H} NMR (125 MHz, CDC13): 137.81, 116.67, 107.78, 104.30, 39.77, 37.123, 25.91, 22.74. Anal. Found (Calc.) C: 78.21 (78.46); H: 8.91 (8.47); N: 12.83 (13.07). 117 Synthesis of l,l-bis(a-pyrrolyl)-3,3,5,5-tetramethylcyclohexane (H2tmcpm) H n \ I \ / An oven dried 100 mL round bottom flask was charged with 3,3,5,5- tetramethylcyclohexanone (1.0 g, 6.5 mmol) and pyrrole (9.3 g, 138 mmol) and capped with a septum. The solution was then degassed with argon for 10 min. Trifluoroacetic acid (0.074 g, 0.65 mmol) was added via syringe. The reaction mixture was allowed to stir for 15 min under an atmosphere of argon before being quenched with 0.1 M NaOH (30 mL) solution. The solution was then transferred to a separatory funnel and was extracted with OEt2. The aqueous layer was washed with OEt2 (2 X 30 mL). The combined organic layers were dried with MgSO4 and subjected to rotary evaporation to yield a viscous brown oil. The excess pyrrole was removed by distillation under vacuum (~1 torr). The product was purified by column chromatography on silica gel with an eluant of hexaneszethyl acetate (7:3) to yield a white solid, 0.426 g (24%). M. p. 83-85 °C. 1H NMR (CDC13, 500 MHz): 7.65 (br s, 2 H), 6.55 (dd, 2 H, JHH = 2.5 Hz, JHH =1.5 Hz), 6.13-6.07 (m, 4 H), 2.00 (s, 4 H), 1.27 (s, 2 H), 0.94 (s, 12 H). 13C {'H} NMR (CDCI3 125 MHz): 138.7, 116.4, 107.7, 104.0, 51.9, 47.9, 39.5, 32.7, 31.7. Anal. Found (Calc.) C: 80.17 (79.95); H: 9.94 (9.69); N: 10.46 (10.36). 118 Synthesis of Ti(NMe2)2(cpm) (24) /T'<'NM92 NM92 All manipulations were carried out in an inert atmosphere dry box filled with purified dinitrogen. A 20 mL scintillation vial was loaded with Ti(NMe2)4 (3) (0.222 g., 0.991 mmol) and 2 mL of ether. In a separate vial was loaded H2cpm (0.212 g., 0.991 mmol) in 2.5 mL of ether. The two vials were placed in a liquid nitrogen cooled cold well where they sat until frozen. To a thawing solution, H2cpm was added to Ti(NMe2)4 (3). The solution was allowed to warm to room temperature where it was left to react for 18 h. The volatiles were removed under reduced pressure, which provided the compound in pure form as judge by NMR and elemental analysis. M. p. 126-130 °C (dec). 1H NMR (CDCl3, 500 MHz, 25 °C) 6 7.01 (br s, 2 H), 6.41 (br s, 2 H), 6.30 (br s, 2 H), 3.24 (s, 12 H), 2.19 (br s, 4 H), 1.51 (br s, 6 H). 1H NMR (CDC13, 500 MHz, —40 °C): 7.25 (app 8, 1 H), 6.83 (app s, 1 H), 6.78 (app s, 1 H), 6.73 (app 8, 1 H), 6.11 (S, 1 H), 5.93 (S, 1 H), 3.34 (s, 6 H), 3.15 (s, 6 H) 2.76 (d, 1 H, JHH = 12 Hz), 2.29 (d, 1 H, JHH = 14 Hz), 1.87 (app t, 1 H, JHH = 15 Hz), 1.79 (app t, ‘1 H, JHH = 8.5 Hz), 1.56-1.72 (m, 3 H), 1.42-1.56 (m, 1 H), 1.28-1.40 (m, 1 H), 1.18-1.26 (m, 1 H). 13C NMR {1H} (CDC13, 125 MHz, —40 °C): 163.35, 160.79, 126.54, 123.97, 118.13, 115.33, 106.27, 100.67, 48.12, 47.32, 43.93, 38.81, 37.62, 25.60, 23.37, 23.26. Anal. Found (Calcd.) C: 61.91 (62.07); H: 8.49 (8.10); N: 15.62 (16.08). 119 Synthesis of Ti(NMe2)2(tmcpm) (25) All manipulations were carried out in an inert atmosphere dry box filled with purified dinitrogen. In a 20 mL scintillation vial was loaded Ti(NMe2)4 (3) (0.195 g, 0.869 mmol) and 2 mL of ether. In a separate vial was loaded H2tmcpm (0.270 g, 0.870 mmol) and 2.5 mL of ether. The two vials were placed in the cold well where they sat until frozen. To a thawing solution of Ti(NMe2)4 (3) was added H2tmcpm. The solution was allowed to warm to room temperature, where it was left to react for 18 h. The volatiles were removed under reduced pressure, which provided the compound in pure form as judged by NMR and-elemental analysis. M. p. 174-177 °C (dec). 1H NMR (CDC13, 500 MHz, 25 °C): 6 6.97 (br s, 2 H), 6.37 (br s, 4 H), 3.24 (br s, 12 H), 2.6-1.6 (br s, 4 H), 1.25 (s, 2 H), 0.92 (br s, 12 H). 1H NMR (CDC13, 500 MHz, —40 °C): 7.19 (app s, 1 H), 6.77-6.75 (m, 3 H), 6.11 (t, 1 H, JHH = 2.6 Hz), 5.88 (dd, 1 H, JHH = 1.83 Hz, JHH = 1.09 Hz), 3.35 (S, 6 H), 3.12 (s, 6 H), 2.85 (d, 1 H, JHH = 14.0 Hz), 2.29 (d, 1 H, JHH = 15.0 Hz), 1.73 (d, 1 H, JHH = 15.0 Hz), 1.63 (d, 1 H, JHH = 14.0 Hz), 1.23 (br s, 2 H), 1.00 (s, 3 H), 0.93 (s, 3 H), 0.90 (s, 3 H), 0.63 (s, 3 H). ”C {'H} NMR (CDC13, 125 MHz) (—40 °C): 166.21, 162.59, 126.09, 123.93, 117.15, 115.16, 106.31, 101.06, 51.89, 50.05, 49.45, 48.16, 47.25, 43.43, 37.10, 35.82, 32.14, 31.91, 29.97, 27.40. Anal. Found (Calcd.) C: 65.13 (65.34); H: 8.91 (9.23); N: 13.62 (13.85). 120 u— 1' 5.5 References H . Y. Shi, C. Hall, J. T. Ciszewski, C. Cao, A. L. Odom Chem. Commun. 2003, 586. 2. For some additional reports of dipyrrolylmethane ligands on transition metals see (a) A. Novak, A. J. Blake, C. Wilson, J. B. Love Chem. Commun., 2002, 2796 (b) S. Majumder, A. L. Odom Organometallics, 2008, 1174 (c) J. B. Love, P. A. Salyer, A. S. Bailey, C. Wilson, A. J. Blake, E. S. Davies, D. J. Evans Chem. Commun, 2003, 1390. 3. Majumder, S.; Odom, A.L. submitted. 4. Littler, B. J.; Miller, M. A.; Hung, C. H.; Wagner, R. W.; O’Shea, D. F.; Boyle, P. D.; Lindsey J. S. J. Org. Chem, 1999, 64, 1391. 5. For a recent computational study on pyrrolyl isomerization see Dias, A. R.; Ferreira, A. P.; Veiros, L. F. Comptes Rendus Chimie, 2005, 8, 1444. 6. Swartz II, D. L.; Odom, A. L. Organometallics, 2006, 25, 6125. 7. Tanski, J. M.; Parkin, G. Organometallics 2002, 21 , 5 87. 8. B. K. Carpenter, Determination of Organic Reaction Mechanisms, John Wiley & Sons: New York, 1984. 9. D. C. Bradley, 1. M. Thomas,J. Chem. Soc. 1960, 3859. 10. C. L. Perrin, J. D. Thoburn, J. Kresge, J. J. Am. Chem. Soc. 1992, 114, 8800. 121 CHAPTER 6 Uranium (VI) bis(imido) pyrrolyl complexes: synthesis, structure, and reactivity 6.1 Introduction The metal-ligand multiple bond functional group has played a significant role in transition metal synthetic organometallic chemistry and catalysis.1 This functionality has been far less developed for actinides. Since its discovery, the most widely studied metal ligand multiple bond in actinide chemistry is the uranyl ion, UO22+. The reactivity of the uranyl ion has been generally limited to equitorial coordination sites due to the high degree of thermodynamic stability and kinetic inertness of the U—0 bond.2 Therefore, uranyl analogues are highly attractive to expand on the reactivity associated with metal ligand multiple bonds in actinides as well as investigate f-orbital involvement in bonding. Until the recent work done by Boncella and co-workers in developing a dependable synthetic strategy for uranium bis(imido) analogues,3 the understanding of the extent to which the f-orbitals participate in U-element multiple bonding was limited to the uranyl ion and a hand full of uranium imido complexes.4 Analysis of chemical bonding using density functional theory (DF T) On uranium bis(irnido) complexes suggests that there is a significant amount of covalency in the U—N bond, which goes against the paradigm of the actinide elements being highly ionic in nature.5’6 The 6 orbitals involved in bonding in U(NMe)2I2(THF)2, 0,, 0,, two as and mu, are the same types involved in the uranyl ion, although the ordering differs, which is partly due to the higher electronegativity of the oxygen atom and larger involvement of the 6p orbital in uranyl.“7 122 Unlike uranyl, the U—N multiple bond in U(NBut)2(THF)2I2 is not kinetically inert. The U—N bond can undergo [2+2]-cycloaddition with aryl isocyanates to yield an array of uranium (bis)imido derivatives,8 as well react with B(C6F5)3°(H2O) to yield an uranium oxo imido complex (Scheme 6.1).9 Interestingly, the addition of OPPh3 forces the iodide ligands trans. Scheme 6.1 Exchange of imido ligand in U(NBut)2(THF)2I2 and t U(NBu )2(OPPh3)2lz. \rL/ O THF\||/| 3(Cerla' ”20 THRfl/I OPPh, Ph3P°;U\/' /U\ ’ THF/“\I ’ ' || OPPhs THF || 1 N A A Ar i ll Ph3PO\ /1 Ph3P0\ ,1 U ArN=C=O l/“\OPPh3 e l/||\0PPh3 2, )1 Among the many ancillary ligands employed in organoactinide chemistry, the most common are Cp-based ligands. Burns and Amey reported a cis-imido uranium complex U(115-C5Me5)2(NPh)2,4c while Boncella and co-workers reported U(ns- C5H5)2(NBut)2 with the irnidos in a cis-geometry. They also prepared a mono-Cp based system in U(NBu’)2(dmpe)l(n“-C,H,) with the imido ligands trans.” Similar to Cp, pyrrolyl ligands can adopt bonding modes of n5 or n' (Chart 6.1). Marks and co-workers reported a tetrakis(pyrrolyl) uranium complex where 3 of the 123 pyrrolyls are nl-bound and the other is 115-bound.ll The fluxional behavior of pyrroles involving interchanging n'- and 115-coordination in their bonding hapticities is common in transition metals. Similarly at elevated temperatures the pyrrolyls interchange rapidly on the NMR timescale between 715- and nl-bound in Marks’ uranium pyrrolyl complex.ll Our group has reported titanium ligated dipyrrolylmethanes as effective catalysts for hydroamination and multi-component coupling reactions.12 Pyrrolyl ligands have proven to be a useful class of ancillary ligands for transition metals,13 whereas their employment in actinide chemistry is relatively scarce and to the best of my knowledge there are no reports of dipyrrolylmethane ligands on uranium. I wanted to investigate the scope of dipyrolyhnethane ligands as suitable ancillary ligands for uranium. This chapter discusses the synthesis, structure, and reactivity of uranium bis(imido) dipyrrolylmethane complexes. Chart 6.1 Possible bonding modes of pyrrolyl and dipyrrolylmethane ligands to one metal center. 124 6.2 Uranium bis(imido) dipyrrolylmethane complexes Encouraged by our previous results of pyrrolyl ligands supporting metal ligand multiple bonds on transition metals I sought to prepare novel uranium bis(imido) dipyrrolyhnethane complexes. 1 was intrigued to investigate the bonding modes the ligand may adopt given the reports of Parkin and Tanski and our own research group that increased sterics lower the barrier for pyrrolyl exchanged”l4 Therefore, I thought to use a 2,2’-di(aryl)dipyrrolylmethane (H2dpmmes), where dpmmes is 2,2’-bis(mesityl)-5,5- dimethyldipyrrolyhnethane. The ligation of the dpmme:S on titanium shows pyrrolyls bound-111,115 in the solid state!” While the pyrrolyl groups in Ti(NMe2)2(dpmmes) (13) are inequivalent in the solid-state, the 1H NMR is indicative of fast pyrrolyl exchange on the NMR timescale!“ I was interested to see if dpmmes would exhibit this same fluxional behavior in uranium bis(imido) complexes and what ground-state geometry it would have in the solid-state. Given the wonderful synthetic utility of U(NBut)2(THF)2I2 in preparing a plethora of uranium bis(irnido) derivatives, it was the practical synthon of choice in preparing dipyrrolylmethane uranium bis(imido) complexes. The addition of K2dpmmes to a stining orange solution of U(NBut)2(THF)2I2 in THF provides a black solution, from which U(NBut)2(dpmmeS) (26) can be isolated as a black solid (Equation 6.1). Single crystals of 26 grown from hexane at -—35 °C were suitable for X-ray diffraction. The structure of 26 is shown in Figure 6.1. 125 1 o 0 I e THF H I + / NK KN \ THF, 25 °C,12h / \ 600/0 U(NBu'laldpmmesl 1261 Complex 26 has C2-symmetry with the pyrrolyls bound-715,1)5 in the solid state. The ability of the dpmmes ligand to adopt an n5,n5-binding mode may be partly due to the sheer size of the uranium atom relative to the Size of transition metals. The geometry of the dpmmes ligand forces the imido ligands cis with U-N bond lengths of 1.930(4) A and 1.946(4) A respectively. The average U-N(imido) bond length is significantly longer than the bond lengths found in U(NBut)2(THF)2I2 which is probably due to an electronic effect.3 The average U-N(imido) bond length is similar to the bond lengths found in 115,115-bound U(NBut)2(C5H5)2, although the N-U-N imido angle in 26 is 116.06(15)°, which is significantly larger than the N-U-N imido bond angle CS in U(NBut)2(C5H5)2 (N-U-N = 103.4(3)°), This can be attributed to the chelated dpmm ligand.” 126 .3‘.‘ '2.- t}; 5' ‘ a T” (‘I‘ \ \ ‘\\- ’ r \V t ' at .. ;- f-J '. u . I “ U ' 1. a l ‘ I ‘3 . ‘ v ’5 r--' I. .' Q . o "‘1 _, ' I D 1 J ~ 11121 ’1 -_ "‘ V -' ‘ _ - ‘ ' N ("t r‘ " 4- .1... Figure 6.1 ORTEP structure of U(NBut)2(dpmmes) (26) from single crystal X-ray diffraction. Selected bond distances (A) and angles (deg): U-N(1) 1.930(4), U-N(2) 1.946(4), U-N(3) 2.575(4), U-N(4) 2.592(4), N(1)-U-N(2) 116.06(15), N(1)-U-N(3) 91.90(14), N(1)-U-N(4) 131.47(13), N(2)-U-N(3) 130.79(13), N(2)-U-N(4) 94.11(14), N(3)-U-N(4) 95.72(12). Uranium imidos are known to undergo [2+2]-cycloaddition to yield imido dervatives as well as react with B(C6F5)3'H2O to produce an oxo imido uranium speciess’l0 Attempts to take advantage of this known reactivity with previously reported uranium imidos, showed disappointing imido reactivity in 26, usually resulting in decomposition. Complex 26 is also sensitive to solvent. Leaving 26 in a solution of THF for extended periods of time resulted in decomposition as well, resulting in protio-ligand and new uranium resonances in the 1H NMR spectrum. Unfortunately, the only solvent to CS facilitate the production of 26 from U(NBut)2(THF)2(I)2 and K2dpmm was THF, therefore particular attention to reaction times was required. Attempts to isolate the 127 uranium decomposition product were unsuccessful. To probe the coordination chemistry of 26, a handful of Lewis bases were examined (i.e. pyridine, OPPh3, and dmpe). Most bases examined yielded no clean isolable products, however a reaction of 26 with dmpe, (dmpe = 1,2- biS(dimethylphosphino)ethane), in toluene yields U(NBu’)2(dpm"'°s)(dmpe) (27) (Equation 6.2). A structure for 27 is shown in Figure 6.2. The addition of dmpe to 26 forces the imidos trans, and the increased sterics around the metal center forces the dpmm“ ligand to adopt an n'ml-binding mode resulting in 27 taking on a pseudo- octahedral geometry. Figure 6.2 ORTEP structure of U(NBu’)2(dpm‘“es)(dmpc) (27) from single crystal X-ray diffraction. Selected bond distances (A) and angles (deg): U-N(1) 1.865(15), U-N(2) 1.857(14), U-N(3) 2.393(15), U-N(4) 2.431(15), U-P(l) 3.043(6), U- p(2) 3.116(6), N(1)-U-N(2) 165.9(6), N(1)-U-N(3) 9710(6), N(1)-U-N(4) 98.2(6), N(2)- U-N(3) 92.7(6), N(2)-U-N(4) 91.4(6), N(3)-U-N(4) 93.1(5), N(2)-U-P(1) 84.1(5), N(1)- 128 U-P(1) 84.5(5), N(3)-U-P(1) = 169.4(4), N(4)-U-P(1) = 97.0(4), N(2)-U—P(2) = 80.6(5), N(1)-U-P(2) 87.3(5), N(3)-U-P(2) 102.2(4), N(4)-U-P(2) 163.0(4), P(1)-U-P(2) 67.39(15). The average imido bond length in 27 is 1.861 A, which is Similar to the average bond lengths found in U(NR)2(THF)(I)2.5 The N-U-N imido bond angle, 165.9(6)°, in 27 deviates from the ideal bond angle of an octahedral complex of 180°, which is likely due to steric effects. The average U-P bond distance of 3.079 A is longer than the bond distance in U(NBut)2(dmpe)(n5-C5H5)2 (U-P average = 2.99 A), the only structurally characterized other uranium (VI) complex supported by the dmpe ligand.10 The 1H NMR spectrum of 27 shows an extremely broad singlet at 3.35 ppm, assigned to the ortho- methyl groups on dpmmes, which could possibly be due to hindered rotation about pyrrolyl-mestiyl bond (See Appendix for the 1H spectrum of 26 and 27); however, fluxional behavior could also explain this phenomena. This broad singlet is also seen in complex 26. \l/ r—\ _ ‘N,..,N’ + —P\ ,P— Tol.,25 °C,18 11v :Pfi‘f: (6.2) 47% 4\ U(NBu')2(dpmm°s)(dmpe) (27) 26 I also investigated the use of the sterically smaller ligand 5,5- dimethyldipyrrolylmethane (H2dpm). The dry addition of K2dpm to a stirred dilute orange stining solution of U(NBut)2(THF)2(I)2 in THF affords U(NBut)2(THF)2(dpm) 129 (28) as a red solid, which can be isolated in 42% using this synthetic procedure (Reaction 6.3). Crystallization of 28 from THF/hexane resulted in the growth of crystals suitable for X-ray diffraction. A structure of 28 is shown if Figure 6.3. The dilute reaction conditions for the production of 28 were required due to the low solubility of K2dpm, which resulted in slightly longer reaction times. i 1,“ H I \ THF>U< + / NK KN \ : N>U