42.... 35d wt. a: rt «.4... CW .. . .s Hwfiwu $k D. m 5.... .32.. 1.2: . an... .. L53... 3 .waf‘ ”V “V n. 1 . h Wu 15¢ ‘ . .a.” : Atathuvav‘u‘ Ostny t... .l.‘ :I. t‘ I'- .\.?,: It: 4 .. s. (autumn l 34:15" «i 1.41.1353}! 19503.! 1v diur‘. . fifty «flung, 5 1.... 1......3. .u. 374.... JP”! 5353‘ 2v :1: in LI DHHn I Tina's-s Michigan State 3 University 202189 This is to certify that the dissertation entitled ASYMMETRIC SYNTHESIS WITH VAPOL DERIVATIVES AND NOVEL CHIRAL THIOUREA ORGANOCATALYSTS presented by KONSTANTINOS RAMPALAKOS has been accepted towards fulfillment of the requirements for the PhD. degree in CHEMISTRY Major Professor’s Signature 01/7/67 Date MSU is an afiinnative-action, equal-opportunity employer ~.-.—.—-.-a—n—A—-----—--n—--4------—.-:-.— - - 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:lProj/Acc8Pres/ClRC/DateDueindd ASYMMETRIC SYNTHESIS WITH VAPOL DERIVATIVES AND NOVEL CHIRAL THIOUREA ORGANOCATALYSTS By Konstantinos Rampalakos A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2008 ABSTRACT ASYMIVIETRIC SYNTHESIS WI'I'H VAPOL DERIVATIVES AND NOVEL CHIRAL THIOUREA ORGANOCATALYSTS By Konstantinos Rampalakos This thesis describes the investigation of the catalytic asymmetric imino—aldol reaction using VAPOL-derived catalysts, and the development of novel chiral thiourea catalysts for asymmetric transformations, specifically for the aza-Henry reaction and for the Michael-type addition of nitroalkanes to nitroolefins. Concerning the first project (Chapters 1 and 2), a derivative of the VAPOL ligand, 7,7’—dimethylVAPOL, was synthesized according to the method developed in our laboratories for vaulted biaryls based on the Snieckus phenol synthesis. The new ligand was designed based on CPK models for the suspected intermediates of the iminoaldol reaction between imines with ketene silyl acetals catalyzed by a Zr-VAPOL catalyst. Attempts to try the new ligand in reactions where VAPOL wasn’t very efficient afforded a significant improvement on the enantioselectivities. These results not only demonstrate the efficiency of the new ligand but also confirm the model that we proposed for the reaction. The rest of the Thesis, and the major amount of work described in this disertation, concerns the design, synthesis and application of novel chiral thiourea catalysts for asymmetric synthesis. Concerning the aza-Henry reaction (Chapters 3 and 4) we evaluated new potential catalysts with two thiourea moieties for simultaneous activation of both reaction components via H-bonding. We identified a new promising catalyst based on BINAM as chiral schaffold. Optimization of the reaction conditions for the aza-Henry reaction led to the development of an efficient protocol for the reaction of N-Boc imines with nitroalkanes. The yields for a variety of substituted imines were modest to good, and the asymmetric inductions were high. To account for the enantioselective reaction of N- Boc imines with nitroalkanes we proposed an intermediate where one thiourea moiety of the catalyst is bound to the imine while the other one is hydrogen bonding to nitromethane. Chapter 5 describes the application of another novel thiourea catalyst to the addition of nitroalkanes to nitroolefins. First, the unasymmetric addition was studied with simple achiral aromatic thioureas as promoters. As a result, we developed the first efficient protocol for diastereoselective synthesis of unasymmetric syn 1,3-dinitro compounds. Subsequently, the asymmetric variant of the reaction was investigated. A novel thiourea/DMAP bifunctional catalyst was made, based again on binaphthyl diamine. The conjugate addition using this catalyst followed a unique selectivity trend according to which, the less catalyst used, the higher the enantioselectivity of the reaction. Investigation of this trend with a series of studies showed that the nitroalkane/catalyst ratio is the determining factor for the selectivities. Optimization of the reaction conditions led to the first highly efficient protocol for the organocatalytic direct conjugate addition of nitroalkanes to nitroolefines with only 2% catalyst loading and ee’s in the mid-90’s for a range of electron deficient and rich nitrostyrenes. Finally, the last chapter (Chapter 6) discusses some efforts towards the synthesis of new biaryl-based thiourea catalysts using vaulted biaryls as schaffolds, in an effort to extend the scope and applicability of the thiourea catalysts that were described in the previous chapters. ACKNOWLEDGMENTS I would like to thank Prof. Wulff for being a great advisor. His broad knowledge and his valuable advice helped a lot in carrying out the reaserch presented herein. It was very exciting to have Prof. Borhan around too. He is a very fun guy and has provided a lot of help and support (despite the centuries-old Greek-Persian conflict). I was lucky to meet great people in the group all these years. Manish Rawat and Billy Mitchell were gret labmates who didn’t only gave valuable help in the beginning, but also made life in the lab very amusing. The same is true for Glenn Phillips as well, and for Zhengsheng Ding. Ding is probably one the funniest guys I’ve met, and also a very good friend. Aman, Victor, Alex, and other people created nice atmosphere to work in. I am also greatfull to my friends outside the lab; Kyoungsoo, Ninneta, Chryssoula, Justas and Kit have been my dearest friends during the last five years, and I will miss their great friendship. Finally, I am greatfull to my family and my beloved fiancée for their love and continous support. iv TABLE OF CONTENTS LIST OF TABLES ................................................................................... vii LIST OF FIGURES .................................................................................... x LIST OF SCHEMES ................................................................................. xii CHAPTER 1 .................................................................................................................... 1 THE CATALYTIC ASYMMETRIC IMINO ALDOL REACTION USING VAPOL- DERIVED CATALYSTS ............................................................................................. l 1.1 INTRODUCTION ............................................................................................... l 1.2 THE IMINO ALDOL REACTION USING Zr-VAPOL CATALYST ................. 2 1.3 CAN A MODIFICATION ON THE CATALYST STRUCTURE EXPAND THE SCOPE OF THE IMINO ALDOL REACTION AND PROBE ITS MECHANISM? 8 CHAPTER 2 .................................................................................................................. 12 SYNTHESIS OF (S)-7,7’-DIMETHYL VAPOL AND EVALUATION OF THE LIGAND IN THE IMINO ALDOL REACTION ........................................................ 12 2.1 SYNTHESIS OF (S)-7,7’-DIMETHYL VAPOL ............................................... 12 2.1.1 Synthesis of 6-methyl-2-phenyl-phenanthrol .................................................. 12 2.1.2 Oxidative coupling of 6-methyl-2-phenyl phenanthrol and deracemization of 7,7’-dimethyl VAPOL. ........................................................................................... 16 2.2 EVALUATION OF 7,7’-DIMETHYL VAPOL IN THE IMINO ALDOL REACTION AND COMPARISON WITH THE PARENT LIGAND ..................... 21 2.2.1 Reactions using the unsubstituted imine 2 ....................................................... 21 2.2.2 Reactions using the ketene acetal 3b ............................................................... 24 2.3 CONCLUSION ................................................................................................. 27 CHAPTER 3 .................................................................................................................. 28 THE CATALYTIC ASYMMETRIC AZA-HENRY REACTION .................................. 28 3.1 INTRODUCTION ............................................................................................. 28 3.2 METAL CATALYZED PROTOCOLS ............................................................. 29 3.3 ORGANOCATALYTIC APPROACHES ......................................................... 33 3.3.1 Chiral Proton Promoted Aza-Henry Reaction ................................................. 33 3.3.2 Cinchona alkaloid-based phase-transfer catalysts ............................................ 34 3.3.3 Chiral thiourea catalysts ................................................................................. 36 3.3.4 Bis-Thiourea Catalysts for the aza-Henry reaction and related transformations43 CHAPTER 4 .................................................................................................................. 45 A NOVEL BIS-THIOUREA ORGANOCATALYST FOR THE ASYMMETRIC AZA- HENRY REACTION ................................................................................................. 45 4.1. IDENTIFICATION OF THE APPROPRIATE BIS-THIOUREA CATALYST AND IMINE SUBSTRATE .................................................................................... 45 4.2. OPTIMIZATION OF REACTION CONDITIONS .......................................... 51 4.3. SYNTHESIS AND EVALUATION OF DERIVATIVES OF CATALYST 52.57 4.4. SUBSTRATE SCOPE ...................................................................................... 64 4.5. AN INSIGHT INTO THE MECHANISM ........................................................ 68 4.5. CONCLUSION ................................................................................................ 74 CHAPTER 5 .................................................................................................................. 75 THE FIRST HIGHLY ENANTIOSELECT IVE ORGANOCATALYTIC DIRECT ADDITION OF NITROALKANES TO NITROOLEFINES ...................................... 75 5.1 . INTRODUCTION ............................................................................................ 75 5.2 PROTOCOLS FOR THE CONJUGATE ADDITION OF NITROALKANES TO NITROOLEFINS .................................................................................................... 76 5.3. THE CONJUGATE ADDITION OF NOTROALKAN ES TO NITROOLEFINES BY MEANS OF THIOUREA CATALYSIS ........................................................... 82 5.3.1. Development of the thiourea catalyzed non-asymmetric addition ................... 84 5.3.2 Development of the asymmetric conjugate addition of nitroalkanes to nitroolefines ............................................................................................................ 95 5.3.2.1 Identification of the optimum chiral thiourea catalyst ................................... 97 5.3.2.2 Preliminary efforts for optimization ........................................................... 103 5.3.2.3 Effect of catalyst loading: discovery and investigation of a highly unusual selectivity trend .................................................................................................... 107 5.3.2.4 Optimization of conditions and substrate scope .......................................... 124 5.3.2.5 Proposed Stereochemichal Model for the asymmetric addition using thiourea 82 ......................................................................................................................... 131 5.4 SUMMARY AND CONCLUSION ................................................................. 134 CHAPTER 6 ................................................................................................................ 135 TOWARDS VAULTED BIARYL-DERIEVED THIOUREAS AND THE SYNTHESIS OF VANAM ............................................................................................................ 135 6.1 Examination of Buchwald amination as a key step in the synthesis of 91 ......... 136 6.2. Examination of a synthesis of 91 through oxidative coupling of arylamines.... 141 6.3. Attempts to synthesize 91 using the Bucherer reaction .................................... 147 6.4. Future Work ................................................................................................... 150 EXPERIMENTAL PROCEDURES ............................................................................. 153 Experimental procedures for Chapter 2 ......................................................................... 153 Experimental Procedures for Chapter 4 ........................................................................ 165 Experimental Procedures for Chapter 5 ........................................................................ 194 Experimental Procedures for Chapter 6 ........................................................................ 209 REFERENCES ............................................................................................................. 214 vi LIST OF TABLES Table 1. Reactions of Imine 2 with ketene acetal 3a using various catalysts ...................... 3 Table 2. Temperature dependence of the asymmetric induction in 10 ............................... 7 Table 3. Small scale oxidative dimerization of 20 and 26 ................................................ 17 Table 4. Examination of various reaction times for the oxidative coupling of 26 ............. 19 Table 5. Deracemization of 7,7’-dimethyl VAPOL ......................................................... 20 Table 6. Comparison of the efficiency of (S)-VAPOL and (S)-7,7’-dimethyl VAPOL in the asymmetric induction in compound 4 .................................................................... 22 Table 7. Temperature dependence on the asymmetric induction of 27 using Zr-VAPOL catalyst ............................................................................................... 25 Table 8. Temperature dependence in the asymmetric induction of 27 using Zr-7,7’-dimethylVAPOL catalyst ..................................................................... 26 Table 9. Palomo’s approach using catalyst 45 ................................................................. 35 Table 10. Takemoto’s Aza-Henry reaction of phosphinoyl imines .................................. 38 Table l 1. Takemoto’s aza—Henry reaction with N-Boc imines ........................................ 39 Table 12. Jacobsen’s Aza-Henry reaction using 49 ......................................................... 41 Table 13. Ricci’s aza-Henry reaction using catalyst 50 ................................................... 42 Table 14. Evaluation of the bis—thioureas in the aza-Henry reaction of N-Boc imine 41 .......................................................................................................... 49 Table 15. Solvent screening for the aza-Henry reaction with catalyst 52 ......................... 52 Table 16. Effect of the base on the reaction .................................................................... 53 Table 17. Optimization of temperature and catalyst loading ............................................ 56 Table 18. Synthesis of thioureas 56a-61a ....................................................................... 58 Table 19. Synthesis of aromatic N-Boc imines 41 ........................................................... 64 vii Table 20. Reaction scope of the aza-Henry reaction of imines 41a-k using catalyst 52 66 Table 21. Aza-Henry reaction of imine 41 using catalyst 66 ........................................... 71 Table 22. Alcantara’s method for the synthesis of 1,3-dinitroalkanes .............................. 77 Table 23. The first asymmetric addition of nitroalkanes to nitroolefines ......................... 78 Table 24. Wang’s protocol using catalyst 74 ................................................................... 80 Table 25. Maruoka’s formal addition using catalyst 75 ................................................... 81 Table 26. The conjugate addition of nitroalkanes to nitrostyrene catalyzed by thioureas .................................................................................................................... 88 Table 27. Optimization of conditions for the reaction using thiourea 47 .......................... 90 Table 28. Reaction scope for the thiourea catalyzed addition of l-nitropropane to B-nitrostyrene ..................................................................................... 92 Table 29. Reactions of nitroalkanes with fl-nitrostyrene using catalysts 52 and 55 .......... 97 Table 30. Wang’s catalyst screening ............................................................................... 99 Table 31. Evaluation of the new catalyst 82 in the conjugate addition ........................... 103 Table 32. Preliminary optimization of conditions using catalyst 82 ............................... 105 Table 33. An unusual selectivity trend observed for the catalyst loading study for 82 109 Table 34. Concentration study of the reaction ............................................................... 114 Table 35. Cat. loading study with stable nitroalkane/catalyst ratio ................................ 116 Table 36. Effect of nitroalkane equivalents ................................................................... 117 Table 37. Survey of additives in the reaction with catalyst 82 ....................................... 121 Table 38. Optimization of reaction variables ................................................................. 125 Table 39. Substrate scope of the asymmetric addition of nitroalkanes to nitrostyrenes using catalyst 82 ...................................................................................... 127 Table 40. Reaction with 2-nitropropane using catalyst 82 ............................................. 130 Table 41. Initial efforts for the synthesis of 96 .............................................................. 138 viii Table 42. Bucherer reactions of VANOL and VAPOL ................................................. 149 ix LIST OF FIGURES Figure l. The VAPOL chiral ligand .................................................................................. 1 Figure 2.1minoaldol reaction between an imine and a ketene silyl acetal .......................... 2 Figure 3. CPK models of Intermediate 7a and of two imines complexed with it ................ 6 Figure 4. Structure of ketene acetal 3b and products 27a, 27b ........................................... 8 Figure 5. CPK models of complexes of VAPOL and 7,7’-dimethyl VAPOL derived catalysts with imines 2 and 8 .............................................................................. 10 Figure 6. Steric strain between imine 2 and the 7-Me of dimethylVAPOL as a possible explanation for the stereochemical outcome of the reactions in Table 6. ......... 24 Figure 7. Design of Takemoto’s catalyst for the activation of the nitro group in the aza-Henry reaction. ........................................................................................................ 37 Figure 8. Jacobsen’s thiourea 49 ..................................................................................... 41 Figure 9. Ricci’s thiourea 50 ........................................................................................... 42 Figure 10. The concept of bis-thiourea catalysis ............................................................. 43 Figure 11. The chiral ligand BINAM ............................................................................. 45 Figure 12. Structures of bis-thiourea catalysts ................................................................. 46 Figure 13. X—ray structure of compound 53 .................................................................... 51 Figure 14. Schreiner’s thiourea 47 .................................................................................. 61 Figure 15. lH-NMR Study as a probe for imine-catalyst interaction ................................ 69 Figure 16. Postulated mechanism for the reaction using 66 with triethylamine ................ 72 Figure 17. Proposed mechanism for Du’s asymmetric addition ....................................... 79 Figure 18. Maruoka’s catalyst for the formal addition to nitroalkenes ............................. 81 Figure 19. The substituted aromatic thioureas used in this work ...................................... 84 Figure 20. Design of a novel bifunctional catalyst based on BINAM ............................ 100 Figure 21. Plot of the asymmetric induction for syn — 803 versus the catalyst loading .. 110 Figure 22. Possible scenarios to account for the catalyst loading effects ....................... 112 Figure 23. Plot of optical purity of 82 versus optical purity of syn-803 ......................... 119 Figure 24. Temperature study of the asymmetric addition ............................................. 123 Figure 25.Postu1ated stereochemical model ................................................................. 132 Figure 26. Vaulted biaryl thioureas and diamines ......................................................... 135 xi LIST OF SCHEMES Scheme 1. Proposed Mechanism for the Imino Aldol Reaction Using Zr-VAPOL catalyst .......................................................................................................... 4 Scheme 2. Synthesis of naphthols through annelation of o—allyl benzamides ................... 12 Scheme 3. A tentative mechanistic rationalization for the anionic benzannelation ........... 13 Scheme 4. The synthesis of VAPOL using an anionic benzannelation as a key step ........ 14 Scheme 5. The Synthesis of 7-methyl-2-phenyl-4-phenanthrol 26 .................................. 15 Scheme 6. Oxidative dimerization of phenanthr0120 for the preparation of VAPOL ...... 17 Scheme 7. The Henry and the Aza-Henry reaction .......................................................... 28 Scheme 8. Useful transformations of fi-nitroamines ....................................................... 29 Scheme 9. Anderson’s aza-Henry reaction ...................................................................... 29 Scheme 10. Shibasaki’s protocol using the heterobimetallic complex 32 ........................ 30 Scheme 11. Jorgensen’s aza-Henry reaction ................................................................... 31 Scheme 12. Palomo’s aza-Henry reaction assisted by Zn(OTf)2 and N-methylephedrine .................................................................................................. 32 Scheme 13. Johnston’s aza-Henry reaction ..................................................................... 33 Scheme 14. Proposed reaction process for the Takemoto’s thiourea-catalyzed aza-Henry reaction ......................................................................................................... 40 Scheme 15. Synthesis of bis-thiourea catalysts 52, 55 ..................................................... 47 Scheme 16. Synthesis of thioureas 53, 54 ....................................................................... 48 Scheme 17. A tentative mechanistic hypothesis that could explain the effect of the base on the reaction .................................................................................................. 54 Scheme 18. Results for the screening of different thiourea derivatives in the aza-Henry reaction at rt ............................................................................................. 60 Scheme 19. A possible mechanism to account for the high induction obtained with 52 62 xii Scheme 20. Possible reaction intermediates .................................................................... 68 Scheme 21. Synthesis of thiourea 66 ............................................................................... 70 Scheme 22. A plausible stereochemical model for catalyst 52 ......................................... 73 Scheme 23. The conjugate addition of nitroalkanes to nitroolefines as a useful route for the construction of valuable functionalities ....................................................... 76 Scheme 24. The conjugate addition of nitroalkanes to nitroolefines promoted by thioureas .................................................................................................................... 83 Scheme 25. Synthesis of aromatic thioureas 47, 76-79 .................................................... 85 Scheme 26. Rate acelaration of the conjugate addition using thioureas 47 , 76, 77 ........... 86 Scheme 27. A possible explanation for the preference of the syn diastereomer formation in the reaction catalyzed by thiourea 47 .......................................................... 94 Scheme 28. Bis-thiourea approach for the conjugate addition ......................................... 95 Scheme 29. Bifunctional thiourea approach for the conjugate addition ........................... 96 Scheme 30. 2-Aminopyridinium catalyst for nitroalkene activation .............................. 10] Scheme 31. Synthesis of catalyst 82 ............................................................................. 102 Scheme 32. Reaction with trans-a -methyl-B -nitrostyrene ............................................. 129 Scheme 33. Synthesis and reaction of the aliphatic nitroolefine 88 ............................... 13] Scheme 34. Buchwald amination of triflate 95 .............................................................. 136 Scheme 35. Proposed synthesis of 93 using Buchwald amination as a key step ............. 137 Scheme 36. Synthesis of VANOL triflate ..................................................................... 137 Scheme 37. An attempt for Buchwald amination of 95 using CSZCO3 ........................... 139 Scheme 38. Synthesis and amination of triflate 100 ...................................................... 140 Scheme 39. Oxidative coupling of 102 using CuCl2 ...................................................... 142 Scheme 40. Postulated mechanism for the CuCl2 mediated oxidative coupling of substituted anilines ..................................................................................... 142 xiii Scheme 41. Attempted oxidative coupling of 101 using CuCl2 ..................................... 144 Scheme 42. Synthesis and attempted oxidative coupling of 111 .................................... 145 Scheme 43. Attempted oxidation under neat conditions ................................................ 146 Scheme 44. The Bucherer reaction for the synthesis of 104 and 102 ............................. 147 Scheme 45 Bucherer reaction of BINOL ...................................................................... 148 Scheme 46. Cho’s synthesis of BINAM derivatives ...................................................... 150 Scheme 47 Thermal rearrangement of 105 for the synthesis of BINAM ........................ 151 Scheme 48. Proposed synthesis of 91 through hydrazide rearrangement ....................... 15] Scheme 49. Proposed synthesis of 130 .......................................................................... 152 xiv HAPTER 1 THE CATALYTIC ASYMMETRIC IMINO ALDOL REACTION USING VAPOL-DERIVED CATALYSTS 1.1 INTRODUCTION The development of stereoselective transformations for the creation of functionalized optically active molecules is one of the ultimate goals in chemistry, and hence the design of highly efficient catalytic asymmetric reactions has been a major focus of organic synthesis during the last years.1 Consequently, the design and synthesis of new effective ligands has been a great challenge in this field. The Cz-symmetric vaulted biaryl ligand VAPOL2 (2,2’-diphenyl-[3,3’-biphenanthrene]-4,4’-diol, Figure 1) developed in our laboratories has proven to be an excellent ligand for various important catalytic asymmetric reactions, such as Diels Alder reactions’, aziridination reactions4 and imino aldol reactionss. Figure 1. The VAPOL chiral ligand HO 0 Ph H0“ I P“ HO ‘ Ph H0 0 1’“ (S)-1 VAPOL (R)—1 VAPOL The catalytic asymmetric imino aldol rection" (Figure 2) is one of the most important carbon-carbon-bond forming reactions since it provides access to chiral B-amino ketones or esters, which are versatile chiral building blocks for the synthesis of many nitrogen containing biologically important compounds including B-arnino acids and lactams.7 In 1997 Kobayashi reported the first asymmetric Manich type reaction between an aldimine and a ketene silyl acetal using substoichiometric amounts of a Zr-BINOL species as catalyst.8 Their method involves a catalyst generated from zirconium(IV) tert- butoxide and two equivalents of (R)-6,6‘-dibromo BINOL (BINOL=1,1'-binaphth-2-ol). In 2001 we developed an unusually robust Zr-VAPOL catalyst5 that could provide even higher selectivities, and and which retained its selectivity over a range of temperatures from 25 to 100°C, resulting in one of the most efficient protocols for the name reaction to date. Figure 2.1minoaldol reaction between an imine and a ketene silyl acetal .Rg R N 3 oms catalyst TMS. _ NH 0 R 2' + >—< 1.2 THE IMINO ALDOL REACTION USING Zr-VAPOL CATALYST Our interest in the synthesis of chiral amines led us to investigate the use of VAPOL-derived catalysts2 (see Figure 1) for this reaction. The catalyst was prepared by reaction of the ligand with 0.5 eq of zirconium tetraalkoxide in the presence of 0.6 eq of N-methyl imidazole at room temperature for one hour. It was found that VAPOL and 6,6’- dibromo BINOL are superior to BINOL at —45°C5. When the reaction was performed at room temperature, the asymmetric induction for VAPOL remained unchanged, while that for dibromo BINOL and BINOL decreased (Table 1). Table 1. Reactions of Imine 2 with ketene acetal 33 using various catalysts 0H OTMS ca, MKFN + >_< “" NH 0 OMe 0” Ph OMe 2 3a 4 entry ligand mol% cat T(°C) t(h) solvent yield% ee% 1 S-VAPOL 20 -45 20 toluene 92 91 2 S-VAPOL 20 25 15 toluene 94 89 3 R-BINOL 20 -45 19 DCM 80 36 4 R-BINOL 20 25 4 DCM 100 28 R—6,6’-Br2- 5 10 -45 19 DCM 87 86 BINOL R-6,6’-Br2- 6 10 25 4 DCM 87 48 BINOL Catalyst prepared from Zr(OiPr)4/i-PrOH, chiral ligand (2.2 eq) and 1.1 eq of N-methyl imidazole (NMI) in DCM or toluene at 25°C for 1 h. Reactions were performed with 1.2 eq of ketene acetal and 0.125M in imine. (entry 2: 0.5M in imine) The mechanism we proposed for the catalytic cycle involves a catalyst bearing two VAPOL ligands on one zirconium and the coordination of the o-hydroxyphenylimine to the zirconium as a bidentate ligands. It is clear from the examination of spacefilling CPK models that it is possible to bind two VAPOL ligands to one zirconium atom only with a facial arrangement of the four oxygen atoms as is illustratred by structure 73 in Scheme 1. This is supported by lH-NMR experiments on a catalyst generated from zirconium tetraisopropoxide and VAPOL in the presence of two equivalents of N-methyl imidiazole. A clean spectrum is only observed with two equivalents of VAPOL relative to zirconium and the spectrum is consistent with a single C2-symmetrical species which is tentatively identified as structure 7b bearing mutually trans NMI ligands bound to the zirconium.“10 An approach of imine 2 to the open apical position in intermediate 73 is proposed to lead to intermediate 5 in which a phenol exchange has occurred. Scheme 1. Proposed Mechanism for the Imino Aldol Reaction Using Zr-VAPOL catalyst Q OTMS OH H— ? OMe N\ 2 Uh ON 3a b (:H ”\WO) NMI ph 5 X s/ 0 Mi O/|\ 0 M30 / HO NMI C‘xzk/ D 7a: X=open V TMS/ Ph\ 7b:X=NMl NMI 0 fi 6 OTMS NH 0 Ph OMe 4 Reaction of species 5 with the ketene acetal would give intermediate 6 and then release of the product would regenerate the unsaturated species 73 and complete the cycle. 4 A space-filling CPK model of intermediate 7a is shown in Figure 3 and illustrates the binding cleft that is available for docking with imine 2. When imine 2 is bound to zirconium un the cleft, CPK models reveal that there is some freedom of movement up and down in the cleft. The figure depicts the imine rotated up and out of the cleft as far as possible. This conformation should lead to reduced selectivity. The model in Figure 3 predicts that the re—face will be the more shielded face of the imine and this is consistent with our observations. The CPK model indicates that imine 8 with a methyl group ortho to the phenol function should be sufficient to push the imine down into the cleft, thus increasing the selectivity of the reaction. Figure 3. CPK models of Intermediate 7a and of two imines complexed with it Intermediate 7a with Intermediate 7a with CH3 A P“ N oI H PhAN CH3 OH 2 8 Indeed, when the ortho position was methylated there was a profound increase in the asymmetric induction.5 The rate of the reaction of imines with ketene acetals with the VAPOL catalyst is slower with imines generated from substituted aminophenols, however, as indicated in Table 2, this effect can be offset by performing the reaction at a higher temperature where greater turnover numbers are observed. The induction with imine 8a (R=phenyl) shows absolutely no temperature dependence over the range of 25 to 100°C. Table 2. Temperature dependence of the asymmetric induction in 10 Qon Q 0 /\ + >=< R \ N OMe OH TMS Zr/(R)-VAPOL cat. NH 0 toluene ’ R/>99 8 3,4-(MeO)2 2 100 85 96 9 l -Naphthy1 2 100 83 93 Reaction time: 2-3 h. (Entry 1: time 24h) Furthermore, the reaction at 100°C can be performed with an order of magnitude change in catalyst loading with no loss in induction. The electronic nature of the imine has a small effect on the induction at this temperature with a slight increase noted for a p- Reaction time 15 h, entry 5: 5h, entries 6, 7, 8, 9: reaction methoxy substituent and a slight decrease for a p-chloro substituent. 1.3 CAN A MODIFICATION ON THE CATALYST STRUCTURE EXPAND THE SCOPE OF THE IMINO ALDOL REACTION AND PROBE ITS MECHANISM? For imine 2, that bears the non-substituted arene ring, selectivities are lower than for imines 8. Additionally, the corresponding reactions with the unsubstituted ketene acetal 3b derived from t-butyl thioacetate (Figure 4) mediated by the same catalyst are slower and do not display the same temperature independences. For example, it was reported that the reaction of imine 8 when R=1-naphthyl with 3b gives 27b in 91% enantiomeric excess at 25°C (20 mol% catalyst) and in 76% enantiomeric excess at 100°C (5 mol% catalyst). Similarly, the reaction of imine 8 when R=Ph with 3a gives 27a in 71% ee at 25°C (20 mol% catalyst) and 60% ee at 100°C (2 mol% catalyst) .5 Figure 4. Structure of ketene acetal 3b and products 27a, 27b. (1“ filo“ OTMS =< NH 0 NH O t , t S Bu (DA/“\Stau O S Bu 27a 0 27b 3b To improve the selectivities of these reactions, a slight modification on the catalyst structure was envisaged. As it was described aboved (Figure 3), the source of selectivity is the shielding of the re-face of the imine by the phenanthrene unit of the ligand, so that the si-face is exposed for attack by the ketene silyl acetal. This model predicts that any modification that would make the phenanthrene unit bulkier would result in more efficient shelding and thus in higher selectivities. Accordingly, a methyl group on the 7 position of the phenanthrene skeleton would increase the size of the phenanthrene unit and serve to test this model. Figure 5 shows the space-filling CPK model of the complexes between the 8 imines 2 and 8 and the 7,7’-dimethylVAPOL (11) catalyst. The CPK model suggests that there is a significant increase in the steric shielding provided by the phenenthrol of 7,7’- dimethylVAPOL (7-methyl shown in black-red color) compared to that of VAPOL for both imines 2 and 8, and that such a derivative of VAPOL would be expected to increase the selectivities for the above transformations. Figure 5. CPK models of complexes of VAPOL and 7,7’-dimethy1 VAPOL derived catalysts with imines 2 and 8 .. 0 . so“ 11 (S)-7,7’-dimethyl VAPOL Intermediate 711 with dimethylVAPOL imine 2 catalyst with imine 2 Intermediate 7 a with dimethylVAPOL imine 8 catalyst with imine 8 The next chapter details the synthesis of 7,7’-dimethylVAPOL 11 and its evaluation in the asymmetric imino aldol reaction. Specifically, a comparison between VAPOL and 7,7’-dimethy1VAPOL is described for the reactions between imine 2 and acetal 3a, and between imines 8 and a ketene acetal 3b derived from t-butyl thioacetate. The ultimate purpose of this project is on one hand the expansion of the scope and synthetic utility of these reactions and on the other hand the justification of our model and the proposed mechanism for the reaction. 11 CHAPTER 2 SYNTHESIS OF (S)-7,7’-DIMETHYL VAPOL AND EVALUATION OF THE LIGAND IN THE IMINO ALDOL REACTION 2.1 SYNTHESIS OF (S)—7,7’-DIMETHYL VAPOL 2.1.1 Synthesis of 6-methyl-Z-phenyl-phenanthrol In 1986, Snieckus and his coworkers reported that o-allylbenzamides undewent MeLi induced cyclization to give l-naphthols". o—Allylbenzamides were prepared by the directed ortho-metallation of tertiary benzamides followed by the reactions with allyl bromide12 (Scheme 2). This chemistry has subsequently been used to prepare various substituted naphtholsl 3. Scheme 2. Synthesis of naphthols through annelation of o-allyl benzamides NR2 1. s-BuLi NR2 OH 2- MgBrz MeLi O 3. allyl bromide _ O : DC 71% \ 86% 12 13 14 The precursor 13 of the annelation product was synthesized using the extensively reported directed ortho metallation methodology developed by Snieckus’ group. According to that method —- which is focused on tertiary amides as effective ortho metallation directors” — lithiation of aromatic benzamide 12 is directed ortho to the amide group and subsequent reaction with MgBr2 affords an arylmagnesium bromide species as the transmetallation reaction product. The lithium to magnesium transmetallation in the 12 first step of the sequence is necessary for the successful ortho allylation, since the lithiated species failed to give the ortho substituted product”. A mechanistic rational for the transformation is illustrated in Scheme 3. Formation of the naphthol 14 may occur via intermediate 15 by two pathways. The first involves direct cyclization from anion 16, while the second involves formation of benzocyclobutane 17 followed by a [1,3]-sigmatropic rearrangement”. Scheme 3. A tentative mechanistic rationalization for the anionic benzannelation. NR2 LIO E t o, ,\ 2 7’: ‘—-_- / Li + 16 17 +Li —o NH2 15 A few years ago, Dr Su Yu from our group developed a new strategy for the synthesis of VAPOL that was inspired by the Snieckus’ naphthol synthesis. The synthesis of 2-pheny1-4-phenanthrol 20 (the precursor to VAPOL) was realized from a MeLi induced cyclization of diisopropyl 6-methyl-2-(2-phenylallyl)naphthamide 19, which was obtained by a directed o-allylation of 1-naphthamide 18 (Scheme 4). This unprecedented synthesis of a 4-phenanthrol using the Snieckus method established a new efficient method for the synthesis of vaulted biphenanthrols. 13 The method Dr. Yu developed has a few modifications from Snieckus’ original protocol. For the synthesis of naphthols, Snieckus used diethyl-benzamide, TMEDA as an additive, and commercially available MgBrz. Dr Su Yu found that, for the synthesis of the 4—phenanthrol 20, diisopropyl naphthamide works better than diethyl naphthamide, TMEDA is not necessary, and MgBr2 has to be prepared in situ in order for the transformation to be most efficient. The in situ preparation of MgBr2 can be achieved by reacting Mg turnings with 1,2-dibromoethane in EtzO. In this way higher and more reproducible yields of compound 19 can be obtained. Scheme 4. The synthesis of VAPOL using an anionic benzannelation as a key step 0 N1iPr>2 1)s-BuLi o N(iPr)2 2) Mg/C2H4Br2 0 00 3) Bromo Methyl Styrene A 00 75-80% 19 .. .. e MeLi, THF air HO Ph 00 Ho P, 0 2° 00 1 18 Su Yu’s method for the synthesis of 2-phenyl-4-phenanthrol 20 was applied to the synthesis of 7-Methyl-2-phenyl-4-phenanthrol 26, the precursor to 7,7’-dimethyl VAPOL. The synthesis of this new phenanthrol is outlined below (Scheme 5). l4 Scheme 5. The Synthesis of 7-methyl-2-phenyl-4-phenanthrol 26 COOH o SOCI2 7 Benzene, DMF 0O Me 00 M. 21 ” 5 /° 22 23 O N iPr) , . ( 2 1) s-BuLI o N(iPr)2 (IPr)2NH. EtsN _ 2) Mg/C2H4Br2 O CH20|2, 0°C 00 3) Bromo Methyl Styrene , 00 74 % from 22 Me 24 ”75 % Me 25 “0 Ph MeLi, THF DO 80 % Me 26 The first step involves a formal Diels Alder reaction between toluene and furoic acid to furnish 6-methylnaphthoic acid 22. The product was obtained in ~5% yield. The majority of the mass balance for this reaction (as was described in the original protocol by Price et a1.) is a mixture of high molecular weight acids of complex structure, formed by the condensation of two or three molecules of toluene with one of furoic acid”. The larger acids could be separated by extraction of the reaction mixture with aqueous sodium bicarbonate (the high molecular weigh acids could not be extracted with this method). The 6-methylnaphthoic acid was obtained from the resulting mixture with barium hydroxide, since the barium naphthoate of 22 (and furoate) dissolved in hot water while the high 15 molecular weigh acids formed an insoluble barium salt. Finally, after crystallization from ethyl acetate, 8 grams of quite pure product were available for the next step. Compound 22 was then treated with 1.1 eq of SOCl2 and 0.44 eq of DMF in benzene to afford 6-methylnaphthoyl chloride 23. Assuming that the transformation was quantitative, the crude product was taken on to the next step, where it was treated with 1.2 eq of (i—Pr)2NH and 1.2 eq of RN in DCM to give diisopropyl-6-methyl naphthamide 24 in 74% yield from the naphthoic acid, which is consistent with the parent ligand synthesis. The stage was set for the key step of the sequence, the ortho-allylation / cyclization reactions that would give the desired phenanthrol. Accordingly, following the modified procedure of Dr. Yu, the 6-methyl naphthamide 24 was treated with 1.1 eq of s- BuLi, 3 eq of freshly prepared MgBr2 and 2 eq of a-bromomethyl styrene to furnish diisopropyl 6-methyl-2-(2-phenylallyl)naphthamide 25 in 75% yield. Finally, the cyclization of the o-allyl naphthamide 25 proceeded smoothly upon treatment with MeLi in THF at —78°C to afford the desired 7-methyl-2-phenyl-4- phenanthrol 26 in 80 % yield. 2.1.2 Oxidative coupling of 6-methyl-2-phenyl phenanthrol and deracemization of 7,7’-dimethyl VAPOL. In 1996 we showed that racemic VAPOL can be prepared efficiently on a 10 g scale by oxidative dimerization of 2-phenyl-4-phenanthrol 20, simply by heating the neat solid at 190°C in a beaker for ~24h, provided that efficient stirring is supplied in order to maintain sufficient contact of the solid with air). With this method, up to 89% yield of 16 VAPOL can be obtained after a simple extraction of the reaction mixture with ethyl acetate (Scheme 6). Scheme 6. Oxidative dimerization of phenanthrol 20 for the preparation of VAPOL 2° (2333.272, 00 1 Since the synthesis of compound 26 that was described in Scheme 5 afforded only a few hundreds of milligrams of desired product, the oxidative coupling for the synthesis of racemic 7,7’-dimethyl VAPOL had to be performed on a very small scale. Following the procedure described by Bao, oxidative coupling of 2-phenyl-4-phenanthrol 20 and 7- methyl-2-pheny1-4—phenanthrol 26 was examined on small scales (~0.3-0.5 g). The preliminary results are shown below in Table 3. Table 3. Small scale oxidative dimerization of 20 and 26 R O HO Ph 0 O 185-190°C HO ph 00 air HO O P“ R R 00 R scale t (h) yield % H (20) 300 mg 39 48 Me (26) 490 mg 46 33 l7 It was found that the dimerization of 2-phenyl-4-phenanthrol on a 300 mg scale gave a yield of 86% after extraction of the reaction mixture with EtOAc. Although the lH- NMR was clean, TLC analysis showed some minor byproduct spots and baseline material, probably polymeric products. The isolated yield after column chromatography was only 48% (Table 3, entry 1), indicating that the dimerization on 300 mg scale behaves differently than on 10 g scale. This is probably due to insufficient stirring. As the phenanthrol was slowly oxidized to VAPOL, the reaction mixture solidified and finally the stirring stopped due to the high melting point of VAPOL(~250°C) and the small amount of material that was present. In a similar fashion, the dimerization of 7-methyl-2- pheny1-4-phenanthrol gave 33% of 7,7’dimethyl VAPOL after 46 h of heating. Since the oxidative dimerization of the phenanthrols gave a significant amount of byproducts, various reaction times were examined for the reaction of 7-methyl-2-phenyl- 4-phenanthrol in an effort to increase the yield. The results are summarized in the following table. 18 Table 4. Examination of various reaction times for the oxidative coupling of 26 HO Ph I O O 180-190°C HO Ph 26 11 entry t (h) dimer/monomer %yield 11 l 36 7/1 30 2 19 2.7/1 40 3 10 2.3/1 40 4 l 1 1 .5/1 51 5 8 1 .1/1 45 6‘ 10 nd 55 7" 10 nd 67'" 'Reaction performed in an Erlemeyer flask in order to increase the reactive surface. "Reaction repeated using the recovered starting material from entry 6. "'Total yield from entries 6 and 7. It is obvious that by decreasing the time of the reaction, the yield of the desired product increases. It was found that for heating times of >10 h a major byproduct with Rf very similar to that of the desired product was formed (entries 1 and 2) and that compound (not characterized) probably comes from decomposition of 7,7’-dimethyl VAPOL. By decreasing the time of heating to 10-11 h, the decomposition can be prevented, and even if the conversions are lower, the isolated yields are higher. The starting material for these reactions can be recovered and used again. For example, as it is shown in entry 7, an 19 isolated yield of 67% can be obtained based on the recovery of starting material and re- subjection in the reaction conditions. Recently Yu et al. found that the deracemization of VAPOL can be readily achieved in the presence of a copper complex of (-)-sparteine'°. The optimal procedure involves the in situ generation of copper(II) and leads to the formation of (S)-VAPOL in >99% ee from the racemate. Yu’s conditions were used to deracemize 7,7’-dimethyl VAPOL. The results are shown in the table below. Table 5. Deracemization of 7,7’-dimethyl VAPOL O CuCl, (-)—Sparteine 0 Ph HO ultrasound _ HO ph H0 Ph MeOH/DCM H0 P“ l 25°C ‘1 1 1 (3)4 1 entry purity of 11 yield% %ee 1 purified once 65 97.8 2 purified twice' 80 99.0 'A second column chromatography was performed before the reaction to ensure that all yellow impurities were removed. It was apparent from the above results that the purity of the ligand to be deracemized is of crucial importance in order for the reaction to afford enantiopure product (>99%ee). 20 It was also found that the enantioselectivity of the chiral ligand could be further enriched by crystallization from methylene chloride/hexanes. Starting from material of 97.8% ee, the first crop of crystals (~50%) were found to have 99.6% ee, indicating that the (S)-compound is less soluble than the racemic, in contrast to VAPOL. 2.2 EVALUATION OF 7,7’-DIMETHYL VAPOL IN THE IMINO ALDOL REACTION AND COMPARISON WITH THE PARENT LIGAND As was mentioned on chapter 1 (Figure 4), the purpose of synthesizing and studying 7,7’-dimethyl VAPOL in the imino aldol reaction was to test weather the methyl substitution on the 7 and 7’ positions can enhance the asymmetric induction of substrates that didn’t afford excellent selectivities when VAPOL was used as a ligand. The reactions that were chosen as models for the comparison were: (i) Reactions of the ketene acetal 3a with the unsubstituted imine 2 derived from 2-aminonaphthol (Table 6), and (ii) Reactions using the acetal 3b derived from t-butyl thioacetate with imines 8 derived from 2-amino-4,6-dimethyl phenol (Tables 7 and 8). 2.2.1 Reactions using the unsubstituted imine 2 With (S)-7,7’-dimethyl VAPOL in hand, experiments for the comparison of the new ligand with (S)-VAPOL were conducted for the reaction of imine 2 using ketene acetal 3a. The results for the comparison of the two ligands at r.t. and at -45°C are summarized Table 6. 21 Table 6. Comparison of the efficiency of (S)-VAPOL and (S)-7,7’-dimethy1 VAPOL in the asymmetric induction in compound 4. (a) CH Q OTMS cat. A + _ -—-—> NH 0 Ph N > (We 0“ Ph OMe 2 3a 4 entry ligand conditions yield% (b) ee% (c) l (S)-VAPOL rt, 0.5M, Toluene 93 89 2 (S)-MeVAPOL rt, 0.5M, Toluened 94 75 3 (S)-MeVAPOL rt, 0.5M, DCM 100 52 4 (S)-VAPOL rt, 0.125M, Toluene 95 89 5 (S)-MeVAPOL rt, 0.125M, Toluene 97 86 6 (S)-VAPOL -45°C, 0.125M, Toluene 92 91 7 (S)-MeVAPOL -45°C, 0.125M, Toluene 90 91 (a) Catalyst prepared from Zr(OiPr),/i-PrOH, chiral ligand (2.2 eq) and 1.1 eq of N-methyl imidazole (NMI) in DCM or toluene at 25°C for l h. Reactions were performed with 1.2 eq of ketene acetal and were run for 24h. (b) Isolated yield after column chromatography. (c) Determined throuh chiral HPLC analysis. ((1) The ligand was not completely soluble for the given conditions. In the first attempt to compare the two ligands at r.t., it was found that 7,7’- dimethyl VAPOL was not completely soluble in toluene when the reaction was performed at 0.5M in imine (entry 2), and presumably that’s the reason for the significant drop in the ee (75%). The reaction in DCM removed the problem of solubility of the new ligand (entry 3) but the reaction was probably too fast (considering that some evaporation of the volatile solvent may took place) to be selective, affording an even bigger decrease in the selectivity (entry 3, 52% ee). 22 It was then realized that high concentrations are not necessary for the reaction to go to completion in toluene and that lower concentrations should allow the new catalyst to dissolve completely, allowing for an accurate comparison with VAPOL. The optimum conditions for the reactions — concerning catalyst solubility and reaction rate — were met when 0.125M imine was used in toluene. The new catalyst was completely soluble in 0.125M toluene, but, as shown in entries 4 and 5, the enantiomeric excess was slightly lower than that obtained using (S)-VAPOL. Finally, when the temperature was decreased to -45°C, the stereochemical outcome of the two ligands was essentially the same (entries 6, 7). This preliminary comparison shows that the new ligand can be as efficient as (S)- VAPOL at low temperatures for this particular reaction, but at room temperature the asymmetric induction was slightly lower. This is contrary to expectations since the design of the ligand anticipated greater shielding of the imine and greater inductions (Chapter 1, Figure 4). One possible explanation for this outcome is that, the methyl substituent at the 7 position of the phenanthrene ring is bulky enough to make close contacts with the imine, preventing it from properly fitting down into the chiral pocket. In other words, steric hindrance between the phenyl group of the imine and the 7-methyl group of the ligand might be responsible for failure of the imine from properly binding with the catalyst with the result that there is an increase in the exposure of the re-face of the C-N double bond to attack by the ketene acetal leading to decreased selectivity. 23 Figure 6. Steric strain between imine 2 and the 7-Me of dimethylVAPOL as a possible explanation for the stereochemical outcome of the reactions in Table 6. steric Q strain(?) If the above arguments are true, then reactions with imine 8 should not encounter this complication, since CPK models show that the methyl substituent ortho to the hydroxyl group of the imine should push the imine down into the chiral pocket of the catalyst. In this way, the o'-methyl group of the imine 8 should be expected to overcome any problem of steric repulsion between the imine and the ligand. Thus, the rections with imine 8 were chosen for evaluation with the new ligand. The ketene acetal 3b was chosen because this substrate gave less than desired results with the VAPOL ligand.5 2.2.2 Reactions using the ketene aceta13b Despite the high efficiency of the Zr-VAPOL catalyst for the imino aldol reactions using ketene acetal 3a (Table 6), the corresponding reactions with the unsubstituted ketene acetal 3b mediated by the same catalyst were slower and did not display the same temperature independence. The reactions with imines 8a and 8b with ketene acetal 3b with the Zr-(S)-VAPOL catalyst have been reported5 and were reexamined in the present work as control experiments and the data in Table 7 are similar to those reported. The imine 8c was also examined since it gave lower induction with ketene acetal 3 (Chapter 1, Table 2). For imine 8a, in contrast wth the excellent enantioselection observed with ketene 24 acetal 3a at room temperature (98% ee, Chapter 1, Table 2) the reaction with 3b affords only 83% ee (entry 1), while a slight drop is observed at 100°C (entry 2, 80% ee). For imines 8b and 8c the enantioselectivites are also not very high at room temperature (88 % ee and 81% ee respectively) and their decrease at high temperatures is even more significant (80 % ee and 66 % ee, entries 4 and 6). Table 7. Temperature dependence on the asymmetric induction of 27 using Zr-VAPOL catalyst (a) Zr-(S)-VAPOL catalyst OH N : J OH OTMS NH O A" (3b) 1 StBu Ar * SBu Toluene 8 27 % s.m. entry series Ar T°(C) % cat. %yield(b) %ee(c) recovered 1 a Ph 25 20 65 23 83 2 a Ph 100 10 70 22 80 3 b l-Naphthyl 25 20 50 35 88 4 b l-Naphthyl 100 10 6O 30 80 5 c 4-Cl-C6H4 25 20 55 33 81 6 c 4-Cl-C6H4 100 10 6O 3O 66 (a) Catalyst prepared from Zr(OiPr)4/i-PrOH, chiral ligand (2.2 eq) and 1.1 eq of N-methyl imidazole (NMI) in toluene at 25°C for 1 h. Reactions were performed with 1.2 eq of ketene acetal and 0.125M in imine and were run for 24 hours. (b) Isolated yield. (c) Determined by HPLC analysis. 25 The corresponding set of reactions using the catalyst obtained from 7,7’- dimethylVAPOL is shown in Table 8. Table 8. Temperature dependence in the asymmetric induction of 27 using Zr—7,7’- dimethylVAPOL catalyst (a) OH Zr-(S)-7,7'-dimethy|VAPOL N catalyst NH O I > Ar) OH OTMS Ar/I*\/U\StBu (3b) 8 SIBU 27 Toluene %yield %ee entry series Ar T°(C) % cat. %s.m. recovered (b) (C) 1 3 Ph 25 20 65 23 90 2 3 Ph 100 10 70 22 84 3 b l-Naphthyl 25 20 49 35 96 4 b l-Naphthyl 100 10 60 30 88 5 c 4-Cl-C6H4 25 20 56 32 91 6 c 4-Cl-C6H4 100 10 60 30 76 (a) See Table 7. (b) Isolated yield based on recovery of the starting material. (c) Determined by HPLC analysis. The very first attempts to evaluate the new ligand in the iminoaldol reaction of ketene acetal 27 showed that there is a significant improvement in the enantioselectivity compared to VAPOL. For imine 8a where Ar is phenyl the ee rises from 83% [(S)- VAPOL] to 90% [(S)-7,7’-dimethyl VAPOL]. When the reaction is performed at 100°C (Table 8, entry 2) the induction drops a little bit (84%), but it is still higher than that afforded with (S)-VAPOL (80%, Table 5, entry 2). For the other imines these 26 improvements were even greater: For the naphthyl-substituted imine 7,7’— dimethylVAPOL affords 8% higher ee than VAPOL both at rt and at 100°C (entries 3 and 4) while when R is 4-chlorophenyl the increase in the induction is 10% higher at both temperatures (entries 5, 6). 2.3 CONCLUSION It is obvious that the catalyst derived from 7,7’-dimethylVAPOL is quite efficient for the imino aldol reaction with regard to the asymmetric inductions observed with both the ketene acetals 3a and 3b. In the first case where imine 2 was used the reason that dimethylVAPOL does not afford higher selectivities than VAPOL is not very clear, although it is suspected that steric strain between the imine and the 7-methyl group of the ligand may destabilize the normal conformation of the imine in the complex. The reactions of imines 8a-c with acetal 3b turned out to be less temperature independent than the corresponding reactions with ketene acetal 33 (Chapter 1, Table 2), even when 7,7’-dimethy1VAPOL was employed (Table 6), nonetheless it is exciting though to observe that a modification on the ligand structure as small as a methyl group can induce a remarkable enhancement in the enantioselectivity of these transformations. Additionally, these results not only demonstrate the efficiency of the ligand but also help to confirm the proposed model for the reaction. The enhancement of selectivities achieved both through the modification of the imine (Chapter 1) and the modification of the catalyst (Chapter 2) agree with our proposed catalyst structure and its mechanism. 27 HAPT ER THE CATALYTIC ASYMMETRIC AZA-HENRY REACTION 3.1 INTRODUCTION The development of stereoselective carbon—carbon bond forming reactions that create contiguous stereogenic centers bearing heteroatom functionality can provide valuable building blocks for organic synthesis. The aza-Henry (nitro-Mannich) reaction, that is, nucleophilic addition of nitroalkanes to imines (Scheme 7), is such a useful carbon—carbon bond-forming process. The resulting B-nitroamines can be transformed into valuable compounds such as vicinal diamines and a-amino acids by reduction'7 and Nef reactions'8 of the nitro moiety (Scheme 8), respectively. Vicinal diamines are of particular interest owing to their broad utility in the synthesis of biologically active natural products and drug candidates, and also more recently, as chiral ligands for asymmetric reactions”. Although a variety of synthetic procedures have been devised to carry out the Mannich reaction enantioselectively, until quite recently the enantioselective aza-Henry reaction was unknown. Therefore, as an unsolved problem, considerable effort has been directed toward the development of catalytic asymmetric aza—Henry reactions over the past several years. Scheme 7. The Henry and the Aza-Henry reaction o R'ANOZ OH J N02 R Lewis acid R or base R' 11 ./\ 11 JNJR R No2 N, R Lewis acid /I\( N02 R or base R R' 28 Scheme 8. Useful transformations of Iii-nitroamines NHBoc NaNOz, ACOH, DMSO NHBOC N102, NaBH4, MeOHA NHBoc Ph COOH 70% Ph/K/ N02 * NH2 95% Ph This chapter is a review of the available methods that have been developed so far for the catalystic asymmetric aza-Henry reaction, from the early metal-catalyzed examples to the organocatalytic protocols that have appeared more recently. 3.2 METAL CATALYZED PROTOCOLS To the best of our knowledge the first example of a stereoselective (although non- asymmetric) aza-Henry reaction appeared in 1998 by Anderson’s group”. The method they reported was an interesting synthesis of 1,2-diamines, through an addition of lithium nitronates to imines that provided B-nitroamines 29. To avoid retro-addition — a common complication in the synthesis of nitroamines - these intermediates were then reduced with SrnIz, and after PMB cleavage the vicinal diamines 31 could be obtained in high yields and with good diastereoselectivities). Scheme 9. Anderson’s aza-Henry reaction 1) BuLi,THF,-78°C ”02 2) BnN=CHPh _ PMB‘NH Smlz PMB~NH Et 3)THF, AcOH Ph/Kr Et THF/MeOH Ph/kr Et -78 to 0°C N02 60 % NH 28 70% 29 30 2 CAN NH : - 2 MeCN/HZO Ph/Y Et (anti / syn = 10 / 1) 9‘5 % NH2 31 29 Anderson’s group also reported preliminary studies in which imines and nitroalkanes were allowed to react in the presence of Bronsted acids and later the reaction was found to be catalyzed by Lewis acids like BF3 and Sc(OTf)32'. Thus, the stage was set for developing the catalytic, asymmetric variants of this transformation-goals addressed and achieved by the research groups led by Shibasaki and Jorgensen. Shibasaki investigated the aza-Henry reaction conducted with the heterobimetallic catalyst 32 (Scheme 10)”. Since 32 shows concomitant Bronsted-basic and Lewis-acidic behavior, both the electrophile and the nucleophile can be activated. Under optimized conditions 34 could be obtained in 90% yield (synzanti >6: 1) with an enantiomeric excess of 80%. Among the disadvantages are the limitation to nitromethane, the restriction to N- phosphinoylimines 33, and the large amount of catalyst required (20 mol%, which corresponds to 40 mol% chiral ligand). Scheme 10. Shibasaki’s protocol using the heterobimetallic complex 32 CIJtBu DC I“ 00 o; 0‘“ o’ Al_‘,Ph NH2‘ Me - . \N O H2,Raney Nl y /I\/C02Et Me T I T Et Cu‘ _O. HN 0 \N 5”” x 39 plus H l NNrR Ph P H _ 40 .. If silylated nitronates are employed, addition of base is not necessary for reaction with imine 37.23“ Owing to the high reactivity of the silylated nitronates, the reaction proceeds uncatalyzed even at —78°C. Reaction temperatures of -100°C were required to achieve high diastereoselectivities (erythrozthreo >25:1) and high enantioselectivities (>95%ee for the erythro product) in this case. The high selectivity of the addition in both cases is attributed to the Zimmerman-Traxler transition state 40, in which both reaction 31 partners are coordinated to the chiral Lewis acid. If one assumes a rapid equilibration of the E and the Z forms of the Copper nitronate, transition state 40 is formed exclusively, which explains the preferred formation of the erythro product. Very recently, one more important metal catalyzed variant of the name reaction has been reported by Palomo’s group featuring the addition of nitromethane to N-Boc protected imines 4124. The breakthrough in this case is that good to excellent enantioselectivities of the desired products 42 can be obtained by exclusive use of commercially available, inexpensive materials such as Zn(OTf)2 and N-methylephedrine. Scheme 12. Palomo’s aza-Henry reaction assisted by Zn(OTf)2 and N-methylephedrine , Boc 30% 2010102. , 300 IN 30% IPrgEtN HN A J + CH3N02 = A )\/No2 ' 45% (-)-NME ' _ o 4‘ 2° C 59-98% 87-99%ee 32 3.3 ORGANOCATALYTIC APPROACHES The rapid development of organic catalysis25 during the last few years has led to the appearance of several reports concerning the promotion of the aza-Henry reaction through the activation of the imine and/or the nitroalkanes through metal-free catalytic systems. The catalyst that have been used so far can be divided into three groups: (i) Chiral proton catalysts, (ii) Cinchona-alkaloid-based phase-transfer catalysts and (iii) chiral thiourea catalysts. 3.3.1 Chiral Proton Promoted Aza-Henry Reaction In 2004 Johnston et a1. developed a chiral bisamidine triflate salt that effects the diastereoselective addition of nitroethane to a range of electron-deficient N-Boc imines”. The catalyst 44 is a protonated chiral diamine based on diaminocyclohexane as the chiral schaffold. By using 44, neither a Bronsted base additive nor preactivation of the nucleophile is necessary for the addition to occur. Scheme 13. Johnston’s aza-Henry reaction 10% l+ OTI NI-I HN NOZSN H 448>ng ,B J CC + /I\‘/NO2 Ar F1 -20°C Ar/Kl/ 41 R 51-69% 43 d.r.=7:1 to 19:1 59'90‘7066 33 The exact nature of the stereochemical determining catalyst-substrate complex was not clear, however, it was clear from the initial experiments that the proton on the quinoline moiety plays a key role in both substrate activation and orientation leading to asymmetric induction. Johnston’s method is unique in demonstrating the use of a chiral proton (a polar ionic hydrogen bond) alone as both the means of activation (function) and control (structure) of absolute and relative stereochemistry. It also constitutes the first example of a metal-free catalytic asymmetric aza-Henry reaction. 3.3.2 Cinchona alkaloid-based phase-transfer catalysts Palomo’s group found recently that R-amidosulfones are appropriate in situ precursors of enolizable aldehyde-derived azomethine compounds in the context of the asymmetric aza—Heny reaction. Thus, a new asymmetric aza-Henry technology was developed with broad substrate scope based on the use of R-amido sulfone substrates 44 and phase transfer catalysis (P’TC)27. Given that in situ generation of imines from R-amido sulfones requires stoichiometric base”, the base-promoted, nonselective background aza- Henry reaction constituted an initial obstacle”. It was postulated that phase transfer conditions using chiral quaternary ammonium salts in combination with a nonsoluble base would render the competitive undesired reaction marginal. Indeed, it was found that catalytic quantities of 45 in combination with CsOHHzO (120-150 mol %) sufficed for the reaction of nitromethane with a variety of R-amido sulfones 44. The enantiomeric excesses are generally high for in-situ generated aryl-substituted imines irrespective of the electronic nature of the aromatic ring: heteroaromatic imines being also tolerated. Most remarkably, the aza-Henry reaction gave enantiomeric excesses regularly above 94% with 34 an array of enolizable aldehyde-derived imines. Linear as well as branched chain alkyl amido sulfones 44 (R=alkyl) gave the corresponding adduct 46 in good yields and enantiomeric excesses in the 94-98% range indicating no significant sensitivity to the size of the alkyl group in 44. Some representative examples are summarized in the following table. Table 9. Palomo’s approach using catalyst 45 45 (12 mol%) NHBoc \ NHBoc F1 Sng-TOI R' /\ N 02 A R N02 44 CsOH.H20 (130%) 40 9' toluene, -50°C, 44 h entry R R’ yield % ee%(d.r.) l ethyl H 80 96 2 iso-propyl H 8 l 95 3 c-hexyl H 77 98 4 Ph H 79 91 5 2-furyl H 72 84 6 Ph Me 88 94(93z7) Almost at the same time, Ricci’s group reported on the same transformation using the same catalyst 45 but employing KOH instead of CsOH.H203°. His method concerns the addition of nitromethane to aliphatic and aromatic azomethine precursors 44 to afford products 46 (R’=H) with very good yields and ee’s (73—99%). 35 3.3.3 Chiral thiourea catalysts Electrophile activation by chiral small molecule H-bond donors has emerged as an important paradigm for enantioselective catalysis, with new applications and developments appearing at a rapidly increasing pace“. Particularly, the simultaneous action of two hydrogen bond donors has proven to be a highly successful strategy, both in enzymes and in synthetic catalyst systems. Several properties of the bifurcated hydrogen bond may contribute to its utility in catalysis. Such interactions benefit from increased strength and directionality relative to a single hydrogen bond. In an analogous manner, two-point binding is an extremely powerful strategy for asymmetric catalysis with metal- centered Lewis acids”. However, while the requirement for multidentate coordination to a chiral Lewis acid often imposes limitations upon substrate structure, in principle any Lewis base is capable of engaging in bifurcated hydrogen bonds. Electrophiles shown thus so far to be activated by double H-bond donors include aldehydes, ketones, esters, imines, N-acyliminium ions, and nitro compounds3 ‘. Concerning the catalysts used in double H-bond activation, the synthesis of chiral thioureas is facilitated from the ready availability of enantiopure chiral building blocks bearing primary aminofunctionalities. Therefore given the excellent general stability, high conformational rigidity33 and Lewis base binding proclivities of thiourea derivatives, it is unsurprising that chiral analogues are rapidly emerging as versatile, functional group tolerant and easily prepared and modified catalyst templates for the promotion of a wide range of synthetically useful asymmetric carbon-carbon bond forming processes”. 36 Along these lines, given the fact that the nitro group and the imine functionality are both capable of participating in H-bonding, it is not surprising that some protocols concerning thiourea-catalyzed aza-Henry reactions have already appeared in the literature. The first protocol for a thiourea catalyzed aza-Henry reaction was published in 2004 by Takemoto”. His design of the catalyst borrowed from Schreiner’s thiourea 473°, which had been established a few years earlier for the promotion of Diels-Alder reactions. Takemoto’s catalyst 48 was based on the electron withdrawing bis(trifluoromethyl substituted) aryl group which ensures the necessary acidity for the thiourea N-H bonds - and thus the capability in effective H-bond donation —— and incorporates a dimethylamine group as a base for the activation of the nucleophile. Specifically, the corresponding nitronate could be produced from the nitroalkane with the bifunctional catalyst via the hydrogen-bonding activation with the thiourea moiety and subsequent deprotonation by the neighboring tertiary amino group Figure 7. Design of Takemoto’s catalyst for the activation of the nitro group in the aza- Henry reaction. CF3 CF3 or:3 . (1 IL FC NJI\N or: i II 0 CF" 3 H H 3 /N\ 47 48 CF3 CF3 Q i 0 CL JSL Q ,3 '3 I? _ \N: H 11 3 /..\ E_ E _ /+\H ,,,,, g E— cho 0: IO + H/FIFR H/ILR 37 Using catalyst 48 at 10% catalyst loading Takemoto was able to promote the aza- Henry reaction between phosphinoyl imines and nitroalkanes with good yields and modest ee’s. The results for various aromatic and heteroaromatic imines are shown in Table 10. Table 10. Takemoto’s Aza-Henry reaction of phosphinoyl imines n 48 (10%) H INJDPh2 DCM, rt HN,Ph2 Ar) '/ N02 Ar Ph R N02 33 10 eq, 34 entry Ar R yield % %ee (dr) 1 Ph H 87 67 2 4-Me-C6H4 H 72 63 3 4-Cl-C6H4 H 76 67 4 2-naph H 78 70 5 2-furyl H 85 76 6 2-pyridyl H 91 68 7 2-thienyl H 57 64 8 trans-styryl H 68 65 9 Ph Me 83 67 (73/27) The same group reported a second protocol in 2006 for the name reaction using the same catalyst, but featuring N-Boc protected imines”. The reactions with N-Boc imines turned out to be much more efficient in terms of yields and enantioselectivities and a number of higher nitroalkanes could be successfully used affording syn products with high diastereoselectivities as well. Representative examples are given in the Table 11. 38 Table 11. Takemoto’s aza-Henry reaction with N-Boc imines jI’Boc N02 48 (10%) HN'BOC Ar + 6 ; /I\rNo2 n DCM, 20°C A’ n 41 43 entry Ar R yield% % ee 1 Ph H 90 94 2 4-CF3C6H4 H 80 98 3 4-Me H 82 93 4 4-MeOC6H4 H 7 l 95 5 l-naphthyl H 85 95 6 3-pyridyl H 89 98 7 Ph Me 92 93(90/ 10) 8 Ph CHzPh 84 97(83/ 17) 9 Ph CHZOBn 80 95(86/14) Takemoto’s protocol with N-Boc imines was the first report of a highly diastereo- and enantioselective aza-Henry reaction with functionalized nitroalkanes to give N- protected imines. To account for the highly stereoselectivity of the reaction, a ternary complex C of catalyst 48, imine 41 and nitronate anion was proposed as a plausible transition state, in which the imine is hydrogen-bonded to the thiourea moiety and the amonium group of 48 is hydrogen bonded to the nitronate anion (Scheme 14). Ternary complex C can be considered to be generated through two different ways. On one hand, thiourea catalyst first activates the nitroalkane by hydrogen-bonding interaction which is followed by intra- or intermolecular deprotonation by the amino group of 48 to generate 39 nitronate complex B. Subsequent binding of the imine to the thiourea moiety in place of the nitronate produces complex C. On the other hand, complex C might be formed by the successive interaction of the imine and the nitroalkane with thiourea catalyst 48 via binding and deprotonation (D to C). In any event, the thiourea moiety of 48 is believed to play a crucial role in activation of N-Boc imine in the nucleophilic addition step or nitroalkane in the deprotonation step. Complex C correctly predicts both the diastereomeric preference for the syn isomer as well as the absolute stereochemistry observed. Scheme 14. Proposed reaction process for the Takemoto’s thiourea-catalyzed aza-Henry reaction it s -. Ar ’N ' -. ,A 2' H Iii/[LE r —> \N E / ‘HWHF Z _ C .15: + (“’0 nitroalkane |+ / H B ”A” A thiourea 48 1 imine S H \L’fis All. RCH2N02 _ Arm/IL”, M ‘1 C.) III III I 3" Boc NVAr R ,N:—'...H-i"N\ HN’ Cir T 0‘5 s‘ 5 Me No O£u H x If I N 2 H 0811 C major product Another chiral thiourea catalyst successfully employed in the context of the aza- Henry transformation is J acobsen’s thiourea 493°. Thiourea 49 can promote the addition of nitroethane to N-Boc imines with good dr ratios, while the yields and enantioselectivities for the resulting nitroamines are excellent. 40 Figure 8. Jacobsen’s thiourea 49 mg. :0 OPiv Table 12. Jacobsen’s Aza—Henry reaction using 49 1"“ 123.1132. M“ Ar + I/ T Ar N02 Ft toluene, 4°C F1 41 (2.5 eq.) 43 entry R yield % d.r. % ee 1 Ph 96 15/1 92 2 4-Cl-C6H4 98 7/ l 95 3 3-Cl-C6H4 85 7/ 1 96 4 2-Cl-C6H4 99 2/ l 93 5 4-Me-C,5H4 90 12/ 1 96 6 2-Me-C6H4 99 9/ l 95 7 4-MeO-C6H4 95 16/ 1 96 8 3-pyridyl 79 7/ l 97 9 2-furyl 95 6/1 93 10 2-naphthyl 91 5/ 1 97 Finally, Ricci’s group has reported aza-Henry reactions of N-Boc imines with nitromethane using the Cinchona alkaloid-derived thiourea 5039. 41 Figure 9. Ricci’s thiourea 50 01:3 8 6 H N HN—/< 50 HN.,_ /O CO Catalyst 50 can promote the reaction of nitromethanes — but not of other higher nitroalkanes — with N-Boc imines affording good yields and modest to excellent enantioselectivities. Table 13. Ricci’s aza-Henry reaction using catalyst 50 £00 .800 IN 50 (10%) ”N J '1‘ M9N02 /I\/N02 Ar Ar toluene, -24°C 41 (5 eq.) 42 entry Ar yield % % ee 1 l-naphthyl 87 88 2* 2-naphthyl 82 94 3 4-Cl-C6H4 77 94 4* 2-Br-C6H4 82 88 5 4-MeO-C6H4 65 82 6 2—thienyl 50 82 7* 2-furyl 58 63 *Reaction at -40°C 42 3.3.4 Bis-Thiourea Catalysts for the aza-Henry reaction and related transformations In the protocols discussed so far, concerning the mechanistic details of the transformations, it is clear that the imine and the nitroalkane are both capable of forming H—bonds with the thiourea catalysts. Jacobsen’s thiourea 49 (Table 12) is not bifunctional and external base is added for the deprotonation of the nitroalkane, so it is not clear weather it is the imine or the nitroalkane that is interacting with the catalyst in the transition state of the reaction. Takemoto’s and Ricci’s thioureas are bifunctional and so they play a more definite role in the deprotonation step, but again the reaction features two components (imine and nitroalkane) both of which are potential H-bond acceptors, while the catalyst structures have only one thiourea moiety and thus it is not clear which substrate is bound to the catalyst. Our interest in thiourea catalyzed aza-Henry reactions led us to consider the possibility of a catalyst that would incorporate two thiourea moieties within its structural framework, and thus provide the potential for the simultaneous activation of both reaction components, the imine and the nitroalkane. Figure 10. The concept of bis-thiourea catalysis S JL 1’ o--::z 2. ’1 Z \ II O+-I L a: chiral scaffold - . deprotonation 77 2D / )=cn --IZ 2 ----:1:2 x40 /<20 43 A catalyst that would have two thiourea moieties should be able to interact with the imine and the nitroalkane at the same time, activating the two components while bringing them together in its chiral pocket. This could potentially result in well-defined orientations and transition states that would promote high enantioselectivities. Additionally, the concept of the bis-thiourea catalysis is interesting because it could be applied to many other transformations that involve activation through H-bonding. The next chapter details on the development of a novel bis— thiourea organocatalyst for the asymmetric aza-Henry reaction. 44 HAPT ER 4 A NOVEL BIS-THIOUREA ORGANOCATALYST FOR THE ASYMMETRIC AZA-HENRY REACTION 4.1. IDENTIFICATION OF THE APPROPRIATE BIS-THIOUREA CATALYST AND IMINE SUBSTRATE As was mentioned at the end of Chapter 3, the aim of the project presented in this chapter is the identification of an efficient aza-Henry catalyst that would bear two thiourea moieties in its structural framework, for the concurrent activation of the imine and the nitroalkane substrates. Concerning the chiral schaffold of the catalyst, we decided to focus our investigation on l,l’-binaphthyl-2,2’-diamine 51 (BINAM, Figure 11), a chiral diamine that is rarely used in organic catalysis. A search for BINAM in Scifinder gives dozens of results for the use of the diamine as a ligand for metal catalysis, but its use in metal-free catalysis and particularly in the field of chiral thioureas, by comparison, has been very rare 4°. Intrigued by the fact that thioureas based on BINAM are largely unexplored, we designed and synthesized the new chiral thioureas 52 to 54. Figure 11. The chiral ligand BINAM 45 Figure 12. Structures of bis-thiourea catalysts CF3 QC 3 8 CO NJLIN CFa 0C HJINII‘ com cow co 8 s (FD-52 CF3 (R)-53 (FD-54 (Po-55 CF, The design of catalyst 52 borrows from Schreiner’s3° and Takemoto’s35 thioureas featuring the electron withdrawing bis-trifluoromethyl phenyl substituent for enhancing the N—H acidity and thus H-bonding capability of the catalyst. Catalyst 53 was designed based on the speculation that two binaphthyl moieties around the reactive site may create a deeper and more efficient chiral pocket. Thiourea 54 has two thioureas and two histidines that could act as bases, so it was designed as a potentially bifunctional acid/base catalyst. The methylenes between the thioureas and the histidines in 54 would allow for the necessary degrees of freedom of rotation so that an ethyl histidine moiety could fold, approaching a nitroalkane molecule that would be bound to the thiourea and thus deprotonating it, promoting its reaction with an imine, bound to the second thiourea. Finally, the bis-thiourea catalyst 55, which is based on cyclohexyl diamine, is a known catalyst that has been reported for an asymmetric Baylis-Hillman reaction“. Compound 55 46 is the only known bis-thiourea catalyst in the literature, and we decided to evaluate its efficiency in the aza-Henry reaction in comparison with the BINAM—derived compounds. The synthesis of compound 52 is straight-forward. Following the known procedure“ for the synthesis of 55 illustrated in Scheme 1, the new thiourea can be obtained from the reaction of BINAM with 3,5—bis(trifluoromethane)phenyl isothiocyanate in an almost quantitative yield after crystallization. Scheme 15. Synthesis of bis-thiourea catalysts 52, 55 2eq. ,CGS F,C 00 Q 00 i G H H “”2 CF, N N CF, MN”? .N N CF, THF,O°Ctort 00‘ H\n/H 96% S (Fi)-51 (R)-52 CF, 2eq. 04$ F,C N” CFa JSL <1 <1“: a” i H NH2 THF, 0°C to rt “TN CF, (Fi)-56 74% s O The synthesis of catalysts 53 and 54 involves a simple two-step sequence through the intermediacy of the BINAM-bis isothiocyanate 57 (Scheme 16). Compound 57 can be obtained in a nearly quantitative yield from the reaction of (R)-BINAM with thiophosgene 47 under basic conditions.” Without any purification, 57 can be subjected to reaction with another (R)-BINAM moiety furnishing 53 (55%), or with histamine, affording 54 (90%). Scheme 16. Synthesis of thioureas 53, 54 (R)- -BlNAM H1“ 2: 3 0O :0 55% (FD-BINAM K2003 I H20 OO 90% histamine 00 i N’VN (FU'57 900/0 N “W .n N N 00 ”If H “I ,> N (Hi-54 The new bis-thioureas were evaluated in the aza-Henry reaction of nitromethane with N-Boc imine 41. This particular imine was chosen as a model substrate since it has been established in the literature as an imine that is amendable to catalysis by urea and thiourea catalysts” (it is a bidentate imine since it features the N—carboxyl moiety that can potentially participate in double H-bonding). The results of the preliminary screening are shown in Table 14. 48 Table 14. Evaluation of the bis-thioureas in the aza-Henry reaction of N-Boc imine 41 (a) N’ 3°C 20mol% catalyst HN .800 I 4; P11) 10 eq MONOZ Ph * toluene, rt N02 41 42 entry catalyst Et3N (eq.) conversion % (b) % ee (0) 1 (R)-52 - 0 _ 2 - 0.2 20 - 3 (R)-52 0.4 100 (55) (d) 74 4 (R)-53 0.4 100 6 5 (e) (R)-54 - 65 5 6 (t) (R)-54 - 60 4 7 (R)-55 0.2 100 (50) (d) 8 (a) Reaction conducted with MeN02 (10 equiv), and several catalysts (20mol%) in toluene at room temperature overnight (except entry 1, which was run for 2 hours) with 0.25M in imine. (b) Determined by lH-NMR. (c) Enantiomeric excess was determined by HPLC analysis of 42 using a chiral column. (d) Isolated yield after column chromatography. (e) Reaction conducted in DMSO. (f) Reaction conducted in MeNOz (120 eq. with respect to the imine). As can be seen from the data in Table 14, catalyst 52 can’t promote the reaction without the presence of a base (entry 1). On the other hand, the base alone (without the presence of a catalyst) gives a very slow reaction (entry 2). On the contrary, when catalyst 52 is used (20 mol%) along with 40% Et3N at room temperature, a quantitative conversion to the nitroamine 42 along with a very promising enantioselectivity (74%, entry 3) can be obtained. The reaction was very fast at rt, taking only ~2h for complete reaction of the imine. The selectivity of the catalyst 52 is superior to that of catalyst 55 (8% ee, entry 7), indicating that the BINAM schaffold is transfering stereochemical communication to the reaction center much more efficiently than its diamino cyclohexane analog. The isolated 49 yields for these reactions were found to be in the range of 50-55%, and the inconsistency with the observed conversions can be attributed to possible decomposition of the sensitive imine during the course of the reaction, along with the tight separation from the catalyst and aromatic impurities presumably coming from decomposed catalyst in the SiO2 column chromatography, in the case of entry 1 (52 decomposes but 55 does not). Catalysts 53 and 54 turned out to be inefficient in inducing any stereocontrol, affording very low ee’s. In the case of 53, the low asymmetric induction was in contrast to our initial anticipation from the assumption that two BINAM moieties would create an efficient chiral pocket for the reaction. The good conversion observed by lH-NMR for this reaction, though, indicates that the catalyst interacts with the substrates, but obviously in a way that lacks stereocontrol. With a view towards understanding the structure of catalyst 53, we determined the X-ray crystal structure of this compound (Figure 13). To our surprise the structure indicated that the preferred conformation of the bis-thiourea is C2 symmetric only with respect to the relationship between the two binphthyls. The skeleton of the molecule folds in a way that the symmetry within each binaphthyl moiety “breaks” (which is in agreement with the NMR data for compound 53, that indicates the presence of 12 aromatic protons and 20 carbons). The crystal structure also shows that the thiourea moieties adopt an s-trans, cis conformation instead of the expected s-cis, cis. The resulting structure is expected to leave a single N-H bond available for coordination to substrates. This N-H bond seems to be located outside the chiral cavity created by the binaphthyl moieties. The second N-H bond is between the two binaphthyls that seem to Jr-stack, and is probably less accessible. If the same conformation is favored in solution, that would explain the 50 inefficiency of catalyst 53 in exerting stereocontrol. On the other hand, the fact that conversion to the product occurred is interesting because it indicates that even a single thiourea N—H bond alone is enough for substrate activation and reaction promotion. Figure 13. X—ray structure of compound 53 On the other hand, catalyst 53 was found to be highly polar and insoluble in most common solvents, so polar solvents such as DMSO and nitromethane had to be used, leading to almost racemic products (enties 5 and 6). From the above study, catalyst 52 was chosen as the only bis-thiourea that was promising for its reaction outcome, and the investigation was then focused on the optimization of conditions. 4.2. OPTIMIZATION OF REACTION CONDITIONS The search for the optimum conditions was based on four parameters: solvent, the effect of the base, temperature and catalyst loadings. 51 Concerning the optimum solvent, it was anticipated that non polar solvents that are not capable of participating in H-bonding with the substrates should promote the reaction more efficiently, given that a solvent that is an H-bond acceptor can interact with the N-H bonds of the thiourea and thereby “block” the reactive site. The results of the solvent search were in agreement with this notion: As can be seen in Table 15, there is a significant drop in the enantioselectivity when switching to THF (28% ee, entry 4), which presumably interacts with the catalyst, which in addition slows down the rate of the conversion. Benzene and dichloromethane gave satisfactory results, but the selectivity in toluene was the optimum. Table 15. Solvent screening for the aza-Henry reaction with catalyst 52 (a) N,Boc 20mol°/o (FD-52‘ HN.Boc I 7 Ph/l 10 sq MeN02 ph . 0.4 sq Et N SOLVENT? rt ”02 41 42 entry solvent conversion% (b) %ee (c) l toluene 100 (55") 74 2 benzene 100 62 3 DCM 82 59 4 THF 52 28 (a) Reaction conducted with MeN02 (10 equiv), catalyst 52 (20%), EN (0.4equiv) and 0.25M in imine using several solvents. Reactions run overnight. (b) Determined by 'H-NMR. (c) Enantiomeric excess was determined by HPLC analysis of 42 using a chiral column. (c) Isolated yield after column chromatography. 52 With toluene as the optimum solvent, the effect of the base was examined in order to determine how the amount of base affects the yield and the selectivity. As can be seen in Table 16, the amount of the base surprisingly has a determining effect on the selectivity: the optimum induction (74% ee) is obtained when the level of Et3N is 40-50%. Adding a whole equivalent of Et3N causes a significant drop in the cc, and the same happens when only 20mol% Et3N is used. Stronger or weaker bases (entries 5, 6) were found to give almost racemic products. Table 16. Effect of the base on the reaction. (a) N.Boc 20mol% (Po-52 HNIBOC Ph/II 10 egAthNOZ r ph . Toluene, rt N02 41 42 entry base (equiv) pKa yield % (b) % ee (0) l Et3N (0.2) 9.0(DMSO) 30 60 2 Et3N (O .4) >> 50 74 3 Et3N (0.5) >> 50 74 4 EN ( l .0) >> 55 67 5 DBU (0.4) 12(H20) 100 (d) 2 6 N-Me-morpholine (0.4) 7.38 35 4 (a) Reaction conducted with MeNO, (10 equiv), catalyst 52 (20mol%) and 0.25M in imine in toluene at room temperature overnight. (b) Isolated yield after column chromatography. (c) Enantiomeric excess was determined by HPLC analysis of 42 using a chiral column. (d) Refers to conversion by lH-NMR. One could rationalize these results if a few assumptions were made about the mechanism of the reaction. As shown in Scheme 17, it is reasonable to expect that 53 nitroalkane moieties bound to the thiourea will be much more acidic than the ones that do not interact. One would expect that the asymmetric induction in 42 will depend on the concentration of species II, if this is the reactive -and selective- species concerning the catalyzed pathway. The concentration of II will have an upper limit that will be determined not only by the amount of catalyst, but by the amount of the base as well. In other words, species I will have a concentration that will depend on the catalyst itself. For instance, 20% catalyst in a reaction of 0.25M substrate would afford a maximum concentration of 0.050M for I assuming saturation in the binding of the substrates to the catalyst. However, in order for 11 to reach the same level of concentration, the right amount of base, or a base with the proper pr, will be needed. Scheme 17. A tentative mechanistic hypothesis that could explain the effect of the base on the reaction S chiral S chiral JL /scaffolg\ JL 3 scaffold S RN N N NR JL / \ JL 1 1 1 1 RN N N NR *1 t' +1 H —~ A .1 .1 l1 5gb : — O N\ R -O\+,O : : 1 v N - CH, .9: T; B——H CH2 OYNVR (pKa=~10) + R' I II ~3- H C NO H C NO 3 T 2 _, 2 = 2 \J B 111 The pKa of bound MeNO2 is expected to be around 10, similar to the pKa of the compound in water, where it is H-bonded by water molecules in a similar fashion. The pKa of unbound MeNO2 is expected to be about 17, as it is in DMSO (where it isn’t H- 54 bonded). Most of the nitroalkane in not bound, since there is only 20 mol% of the thiourea catalyst and a large excess of nitromethane (10 eq). Based on these considerations, a base with a pKa=~10 would be the right choice for promoting the conversion from I to II, avoiding the formation of species 111. This is in accordance with Et3N being the optimum base among the ones that were tried. It seems that DBU, which is a much stronger base compared to Et3N, is able to promote the formation of III - and thus, the background reaction - to a degree enough for causing a significant drop-off in the selectivity. On the other hand, for the results in entries 1 to 4, it seems that 0.4-0.5 eq. triethylamine is the necessary amount for maximizing the concentration of complex II, and that a whole equivalent of the base will start promoting the background reaction, decreasing the stereoinduction in 42. Concerning N-methylmorpholine, it seems that its lower basicity makes it inefficient in deprotonating nitromethane to the desired degree, and in this case the concentration of the “active catalyst” (or of II) is not very high (although one could argue that this particular base might H-bond to the catalyst through its oxygen atom and decrease the ee via catalyst deactivation). Independent of the validity of the above assumptions and mechanism insights, one thing that is very clear from the above set of data is that triethylamine in the amounts of 40-50% is the optimum for the reaction, and thus we decided to proceed with further efforts to optimize the reaction by using 0.4 eq. triethylamine. Next, the reaction was investigated at lower temperatures in an effort to enhance the asymmetric induction of 42. The results are illustrated in Table 17. 55 Table 17. Optimization of temperature and catalyst loadings (a) N. 3°C Xmol% (FD-52 HN’BOC PhJ 10 eq MeNOz; ph . 0.:oelxc/IerE1I3N N02 41 42 entry cat.loading% solvent T°C conversion% (b) % ee (c) l 20 toluene rt 100 (55) (d) 74 2 20 toluene -5 100 82 3 20 toluene -35 83 (55) (d) 86 4 20 toluene -55 50 84 5 20 mesitylene -35 86 84 6 20 Cl-benzene -35 93 84 7 10 toluene ~10 100 (50) (d) 80 8 10 mesitylene -10 90 82 9 10 toluene -35 53 78 10 5 toluene -10 66 (30) (d) 71 (a) Reactions conducted with MeNO2 (10 equiv), catalyst (R)-52, Et3N (0.4equiv) and 0.25M imine in several solvents. Reaction time was ~36h except for entry 1 (2h). (b) Determined by 'H—NMR. (c) Enantiomeric excess was determined by HPLC analysis of 42 using a chiral column. ((1) Isolated yield after column chromatography. The temperature was gradually reduced from rt to -55°C with maintaining the catalyst loading at 20mol%. As can be seen in entries 2 and 3 in Table 17, lowering the tempeature to —5°C and -35°C resulted in a significant enhancement of the enantioselectivity, to 82% and 86% respectively, without significantly affecting the 56 conversion of the imine. The reactivity of the catalyst was still satisfactory at —35°C, affording the same isolated yield for 42 as for rt. The rate of the reaction at -35°C was much slower than at rt (2h), but this could be overcome by using longer reaction times (36h). A further decrease in the temperature caused a drop-off in the reactivity, so —35°C was chosen as the optimum temperature. Other aromatic solvents such as mesitylene or chlorobenzene were almost as efficient at -35°C as toluene (entries 5 and 6). Reducing the catalyst loading to 10mol% still provided the chiral nitroamine with a high degree of enantioselection (80 and 82% ee in toluene and mesitylene respectively at -10°C). From the above data, the conditions in entry 3 (20 mol% catalyst and -35°C) were chosen as the optimum balance between yield and asymmetric induction. It should be noted here that the background reaction was tested at these conditions (using no catalyst and 40% triethylamine) and was found to give a 9% yield of racemic 42. 4.3. SYNTHESIS AND EVALUATION OF DERIVATIVES OF CATALYST 52 Since catalyst 52 is readily obtained from the reaction of BINAM with 3,5-bis- (trifluoromethane) phenyl isothiocyanate, we reasoned that a derivatization of the catalyst with different functionalities on the aromatic substituents of the thiourea would be very interesting, because it would allow for a practical and convenient, based on the large amount of the commercially available isothiocyanates, comparison with the parent catalyst, and it might lead to the identification of other promising catalysts as well. We synthesized a series of differently substituted bis-thiourea derivatives with variation of the electronic nature as well as on the steric bulkiness of the structures. The synthesis of the derivatives is shown in Table 18. 57 Table 18. Synthesis of thioureas 56a-61a (a) 00 00 ”.0112 THF I I , ”in Ar +00 “infir 811/8 3 (NH 51 56-618 56-61b entry Ar products T°C %yield (b) a %yield (b) b 1 Ph 56 rt 50 50 2 p-NO,-C,H4 57 rt 100 0 3 3,5-Clz-C6H, 58 rt 100 0 4 p-MeO-C,H4 59 rt 1 7 50 5 2,6—Me2-C,H4 60 rt 0 0 6 2,6-Me2-C,H4 60 70 0 50 7 l-naphthyl 61 rt 5 55 8 l-naphthyl 61 70 25 30 (a) Reactions run overnight except for entry 6, run for 2 days. (b) Isolated yield after column chromatography. The reactions for the synthesis of compounds 57a and 58a were fast and quantitative, presumably because of the highly reactive electron-withdrawing isothiocyanates employed. Catalyst 58a was obtained as a pure solid from the crude reaction product that needed no further purification. The reaction of the rest of the isothiocyanates turned out to be more complicated though, because the less reactive they were, the greater the amount of the undesired monosubstituted byproducts 56-61b that was obtained. As a result, for compound 61 the ratio of di- to monosubstituted products was 58 the least satisfactory, affording only a trace of 61a. This could be partially overcome by increasing the temperature to 70°C which afforded a sufficient amount of the bis-thiourea for testing (25%, entry 8). The most demanding reaction was the one in entry 5, presumably because of the steric bulk at the 2,6-positions of the electrophile: compound 60a couldn’t be obtained even at elevated temperatures. The new bis-thiourea derivatives were all tested in the aza-Henry reaction of N- Boc imine 41 with nitromethane at rt. The results are illustrated in Scheme 18. 59 Scheme 18. Results for the screening of different thiourea derivatives in the aza-Henry reaction at rt Boc o ’ t l t . Bo JN 20 /0 ca a ys H N c P“ 10 eq MeN02 Fh/Ifi 42 41 0.4 eq E13N N02 Toluene, rt “/1ng 0O “inflow Ow N III/”Q CO H\II/H\I©LOM9 100%(NMR) 36°/ NMR 568 5% ea 59' 0°/ /oe°e( ) N02 00 Hill/C la E G Hill/3 MN N CO"‘”1”©°‘ CO “PO... 2 58a 73%(NMF1) CI 573 100%(NMR) <5%ee 16%ee 5 0(1ng Ohm/Hg s 518 100%(NM11)D 4 %ee As can be seen from Scheme 18, the only thiourea that didn’t afford a good conversion was 593. This is reasonable, since the NH bonds of that thiourea are expected to be the least acidic, and therefore the least efficient for H-bond interaction with the substrates. All the rest of the thioureas afforded good conversions, but the asymmetric inductions of the product were disappointingly poor. One can expect that the H-bonding ability of thioureas 57a and 58a should be similar to that of 52, but the inductions clearly demonstrate that even if there is H-bonding interaction, the communication of stereochemical information is ineffective. This difference might be associated with the bulkiness of the 3,5—bis trifluoromethyl groups and / or the arrangement in space of 52 compared to the rest of the thioureas. A better understanding of this issue will require further work — e.g. theoretical calculations or X—ray analysis of the catalysts. In any case, the necessity of the 3,5—bis trifluoromethyl groups is not surprising on the basis of the literature associated with thiourea catalysis. Indeed, this particular substituent seems to dominate the structures of most of the thiourea catalysts that have been used. For example, Schreiner has studied the acceleration effects of various aromatic and aliphatic thioureas for a variety of non-asymmetric reactions and he has demonstrated Figure 14. Schreiner’s thiourea 47 rigidifying S-H interaction CF, fl CF, F3C N N CF H H 3 47 that catalyst 47 has been the most effective structure tested”. Schreiner has rationalized his results in terms of the importance of entropic effects; specifically it was proposed that 61 the (computationally determined) rotational barrier of catalyst 47 is relatively high due to an attractive interaction between the ortho-hydrogen atoms, which are polarized by the adjacent electron withdrawing substituent, and the Lewis basic sulfur heteroatom (Figure 14). This rigidifying interaction would minimize entropy loss upon binding of the substrate and thus facilitate catalysis“. In the context of the catalysts 56-61a, the argument concerning the entropic effect of the S-H interaction discussed above could still apply (Scheme 19). In a comparison of catalyst 52 with 58a, for instance, Scheme 19. A possible mechanism to account for the high induction obtained with 52 R N) CI 00 Nil,” “““ 02m — DO j‘I\N’I:j\01 —* ~13 .1. I 0’ ”Q 90 O\ NVR C c. c. r 1 11 n z) CF. N H )c S' S ...o R CO JL 3:1 N’ILN’H“ <— / II II CF" , —» ‘5. H H ”‘1' CO OYNVR F,C CF, R 111 w it could be argued that since the chloro- substituent is not as electronegative as the CF3 group, this attractive S-H interaction shouldbe diminished, and the thiourea moiety would 62 be more flexible in rotating and adopting different conformations, presumably less selective ones. On the other hand, for catalyst 52 the S-H interaction should make complex IV more abundant than III, resulting in bidentate binding and thus more well- defined orientations. 63 4.4. SUBSTRATE SCOPE We next turned our attention to examining the scope of the asymmetric aza-Henry reaction for a series of differently substituted N-Boc imines and nitroalkanes. For this purpose several imines were synthesized according to the standard procedure”. Table 19 shows the two-step sequence for obtaining a variety of electron withdrawing, donating or neutral aryl-substituted substrates 41. Table 19. Synthesis of aromatic N-Boc imines 41 HZNBOC Boc l PhSOgNa; 13: K2C03, Na2$O4 : IN,Boc Ar Y,%%%n:fo Ar SOzPh THF reflux Ar 62 41 entry Ar series %yield 62 %yield 41 1 Ph a 100 98 2 p-Cl-C,H4 b 23 81 3 m-Cl-C,H4 c 77 99 4 o-Cl-C6H4 d 72 98 5 p-Br-C,H4 e 31 64 6 p-MeO-C,H4 f 84 77 7 o-MeO-C,H4 g 90 95 8 3-pyridyl h 68 98 9 p-Me-C,H4 i 74 97 10 l-naphthyl j 59 99 Most of the N-Boc imines were oils, and so they were used as crude products without any purification. It should be noted here that N-Boc imines are very sensitive to moisture and their synthesis and workup requires cautious handling. Additionally, it was found that the degree of their purity greatly influences the outcome of the aza—Henry reaction with catalyst 52, especially regarding the enantioselectivities obtained. For example, when the reaction was performed with impure (partially hydrolyzed) imine 41a that contained some benzaldehyde, product 42 was obtained in only 11% ee. 65 The results for the catalytic asymmetric aza-Henry reaction of the imines 41a-k with nitromethane, nitroethane and nitropropane are summarized in Table 20. Table 20. Reaction scope of the aza-Henry reaction of imines 41a-k using catalyst 52 (a) N , Boc ”1:511:11“ .11 ,3... Ar/II 10 sq RCHzNo: Ph ‘ N02 toluene, -35°C R 41 42, 438-k entry Ar R time (h) adduct %yield (b) %ee (c) (d.r.) (d) 1 Ph H 36 42 55 86 2 p-Cl-C,H4 H 15 43a 62 85 3 m-Cl-C,H4 H 15 43b 55 91 4 o-Cl-C,H4 H 15 43c 61 74 5 p-Br-C,H4 H 24 43d 50 78 6 p-MeOC,H4 H 36 43c 50 89 7 o-MeO-C,H4 H 36 43f 40 65 8 3-pyridyl H 22 43g 63 8 1 9 p-Me-C,H4 H 36 43h 48 86 10 l-naphthyl H 36 431 65 85 11 Ph Me 36 43j 59 70 (e) (77/23) 12 Ph Et 36 43k 63 80 (e) (80/20) (a) Reaction conducted with RCH2N02 (10 equiv), catalyst la (20mol%), Et3N (0.4equiv) and 0.25M imine in toluene at -—35°C. (b) Isolated yield. (0) Enantiomeric excess was determined by HPLC analysis of 7a—l using a chiral column. (d) Diastereomeric ratio was determined by HPLC. (e) Major (syn) diastereomer. 66 As can be seen from the data in Table 20, N-Boc imines with either electron donating or withdrawing groups can afford product 43 with moderate to good isolated yields and overall high selectivities. N-Boc imines with chloro substituents seem to be quite reactive, leading to quantitative conversions in short reaction times (15h) and to good isolated yields of the product along with high enantioselectivities, up to 91% ee (entries 2 to 4). Electron donating groups slow down the reaction, but prolonged reaction times can afford satisfactory conversions and yields, along with good asymmetric inductions. The reason for the drop in the selectivity that is observed for the o-MeO substituted imine is not very clear at this point. On the other hand, electron neutral substituents don’t have any effect on the selectivity of the reaction affording ee’s always in the middle 80’s (entries 9,10). Finally, for the reaction of imine 41 with nitroethane a drop in the enantioselectivity is observed (70%ee, entry 11), while for the reaction with nitropropane, the enantio- and diastereoselectivity are still in a synthetically usefull range (entry 12), demonstrating that the asymmetric aza-henry reaction catalyzed by the bis-thiourea 52 is not limited only to nitromethane, but can be applied also to higher nitroalkanes. A disadvantage of the above methodology is that the catalyst can not be recovered, since it partially decomposes during the course of the reaction and during column chromatography. The decomposition material usually has an Rfvalue similar to that of the product of the reaction, making the purification of some of the products tedious. Nevertheless, satisfactory yields in the range of 40—65% can be obtained for the resulting nitroamines. 67 4.5. AN INSIGHT INTO THE MECHANISM A number of possible reaction intermediates can be envisioned for the reaction with catalyst 52. It can be postulated that the two thiourea moieties can interact with two nitro groups (A), or with two imines (B), or with a nitroalkane and an imine at the same time (C), giving rise to three different species that interconvert with each other. These possible intermediates are shown in Scheme 20. Scheme 20. Possible reaction intermediates . hiral chiral S c s JSL schaffold i JL /50hafi0“\ JL RN N/ \N NH F"? 'i‘ '1' '1‘” '1 t' 11 1.1 _ 2' '1 '9 1: 19,311 15,, $5 OYNVR CYNVR Et N—H " +Ii‘ . 3 +3 CH2 CH3 R R' Z + N62CH2 Ng/Fl A Et3NH Oat B R' \ chiral S schaffold S / JL/ \JL '1'.“ 1 1 11* 1.1 t' 1.1 H _‘ + ‘ Et N—H‘P‘I'I’O- RVNYO C If the reaction occurred through intermediate A, the imine would have to be activated through the protonated triethylamine (if the imine interacting with the complex at all) in a “chiral nucleophile” mechanism. Complex B seems less possible, because the nitro group has been documented in the literature to be a very good thiourea coordinator”. It’s also reasonable to suggest that MeNO2 is interacting with the thiourea, since Et,N is not basic enough to completely 68 deprotonate free MeNO2 (pka=17 in DMSO vs 10 in H20). On the basis of that, mechanism B can be ruled out. Finally, complex C describes our initial idea when designing the bis-thiourea catalysts, that is the dual activation of both reaction components. To distinguish between mechanism A and C, we performed 'H-NMR experiments in order to probe the interaction between the imine and the catalyst. Specifically, we prepared NMR samples of the imine alone and of imine / catalyst mixtures at different ratios. The results of this experiment are summarized in Figure 15. Figure 15. lH-NMR Study as a probe for imine-catalyst interaction A . 8.88pm: 3 i 1 1'1 . 1 i 1, 1 l 'F l‘: I 1AA I . I. 1’ I I n‘ F ,1. I A." . \_ 1.-.... . .‘v , I’ I. \ .‘ “~"-\4 ..-.- —. our-w-“ 7'7 . l r , I ~ - . TY ”Tr-”TI”. I I I I I I 9.2 9.0 8.8 8.5 8.4 8.1 8.0 7.8 7.5 74 7.2 7.0 DUI 1: B 11 A 8.74pp1n ,’ 11‘ I'- .’ » I 1' {I 2 I. I.) I 'II A 312. Si .I‘A I , l: 1 TT‘T‘W -1 T 'TT"" 7 "T, ( I ‘T’. T' 92 90 88 86 8.8 82 00 78 76 74 7.2 709“ .3 i l I f 1 C . 1 ' ‘ 17.? 1 A r I ( - I . i , “ ,1 A I,‘ A. ‘I, .A\ . I, g. 1 I}! :1 8.65m)!!! ' , g- . "I. ix 1 , .. '; A , ~ \... x- i 11> - ~41 , A, _ ,,., _‘ . 7,. _. - ,..-.A_. ,, .-. , _ .. 9.2 9.0 I. O i ' A 8 0 6 0.8 8.2 8.0 7.8 7 6 7 4 7.2 7 0 pp A: NMR of imine; B: leq. imine with 1 eq. catalyst 52; C: leq. imine with 2 eq. catalyst 52 69 As can be seen from the data in Figure 14, there is a significant change in the NMR of the imine while adding first one equivalent (B) and then two equivalents (C) of catalyst. The C(=N)H proton of the imine shifts from 8.88 ppm in the free imine to 8.65 ppm in the imine / 2 eq. catalyst mixture, strongly indicating that there is an interaction between the two species. On the basis of this experiment, mechanism A seems to be less possible than C. Subsequently, species C seems to be the actual intermediate of the reaction, or at least expected to contribute more than A or B to the reaction outcome. Additionally, the importance of both the thiourea moieties in the reaction mechanism was probed. Catalyst 66, that bears only one thiourea in its structure, was synthesized and evaluated in the aza-Henry reaction. The catalyst was accessed in 4 steps from BIN AM 51, according to the procedure reported by Wang and coworkers40b (Scheme 21). Scheme 21. Synthesis of thiourea 66 O CO A_:2.__..O' AcOH formaldehyde; CO “k NH2 O"\NH ..xNHz DCM 2 NaBH,CN "‘N\\ 720/0 ACOH THF 51 85% 64 4MHC| 30H CF, 54% 01115‘ 32:9... co :3 ‘32: 66 65 70 Catalyst 66 was evaluated in the reaction of imine 41 with nitromethane in toluene and DCM and the data is shown in Table 21. Table 21. Ala-Henry reaction of imine 41 using catalyst 66' ,Boc N 20mo|°/o 66 , 300 | HN Ph 10 eq MeN02 ph . " N02 41 42 entry solvent Et3N eq. % yield ee% 1 toluene 0 O - 2 DCM O 0 - 3 DCM 0.4 50 4 ‘Reactions run overnight, with 0.25M imine in DCM or toluene. The reaction using catalyst 66 alone without an external base gave no product, presumably because the dimethylamine functionality of 66 is not basic enough to deprotonate nitromethane. When 0.4 eq. of triethylamine was added to the reaction mixture a 50% yield of the product was obtained, but with almost no asymmetric induction. This experiment shows that the presence of both thioureas is essential for the reaction outcome in terms of stereoinduction. One would expect that the intermediate for the reaction with 66 would feature nitromethane H—bonded to the thiourea, activated for deprotonation, but with the imine probably not interacting directly with the thiourea moiety (Figure 16). That would leave the imine without a well defined orientation, resulting in the very low induction that was observed. These observations with catalyst 66 are in accordance with mechanism C in Scheme 20 for the reactions with catalyst 52. 71 Figure 16. Postulated mechanism for the reaction using 66 with triethylamine I R) imine orientation not well defined The absolute configuration of the products 42, 43 was determined by comparison of the HPLC data and the optical rotations of the products with the data reported in the literature. To account for the enantioselectivity and the diastereoselectivity of the transformation we propose the model shown in Scheme 22. The imine can interact with the thiourea adopting two different orienrations, one where the aryl group points to the catalyst (complexes III and IV) and one where it points away from it (I and II). It is proposed that the orientation of the imine with the aryl group away from the catalyst should be more energetically favored, because this way the strain associated with this interaction is minimized. This orientation exposes the si face of the imine towards nucleophilic attack from the activated nitronate, and this is in accordance with the absolute configuration of the stereogenic center bearing the N-Boc group. Concerning the approach of the nitroalkane, it is clear that there are two different orientations as well. Again, the determining factor is probably the minimization of the steric strain of the alkyl chain of the RNO2 resulting from close contacts with the catalyst structure. This hypothesis is in accordance with the data in Table 20, entries 11 and 12, 72 where a slightly higher selectivity is afforded when R=Et (product 43j) compared with R=Me (product 43k). The triethylammonium ion which results from the nitroalkane deprotonation might interact with the nitronate anion with hydrogen bonding, or with both the nitronate ion and the imine, as in Takemoto’s model .35 Scheme 22. A plausible stereochemical model for catalyst 52 HN,Boc Boc NO a OBTAINED Ar . 2 A ENANTIOMER Ag F3C FA VORED (LESS STRAIN) 73 Although the thiourea moieties in catalyst 52 presumably have some freedom of rotation and other structures and models could be also drawn, the model in Scheme 22 is a working hypothesis which can explain the observed selectivities. 4.5. CONCLUSION In conclusion, a novel BINAM-based catalyst for the asymmetric aza-Henry reaction has been developed. The bis-thiourea catalyst probably functions according to a mechanism where both the components of the reaction are activated simultaneously by hydrogen bonding, indicating a mode of action different from that of the existing protocols and demonstrating the potential of bis-thioureas as an interesting class of unexplored catalysts. The catalyst is very easy to make and modify, and affords the aza-Henry products with modest to good yields, overall high enantioselectivities and synthetically useful diastereoselectivities. 74 HAPT ER THE FIRST HIGHLY ENANTIOSELECTIVE ORGANOCATALYTIC DIRECT ADDITION OF NITROALKANES TO NITROOLEFINES 5.1. INTRODUCTION The use of nitroalkanes as reactive nucleophiles42 and the use of nitroalkenes as Michael acceptors43 have attracted significant interest in recent years in carbon-carbon bond formation reactions. Because of the activating effect of the nitro group, as well as its versatility as a masked functionality, nitro compounds have been quite useful in the synthetic arena“? The Michael addition of nitroalkanes to nitroolefins is particularly interesting because the reaction products, 1,3-dinitro compounds (69), can be easily reduced by H2 to 1.3-diamines 7145, a valuable functionality in the synthesis of complex molecules - given the number of natural products or pharmaceuticals that contain nitrogen — or other 1,3- difunctionalized compounds, heterocycles“, and carbohydrate derivatives47 (Scheme 23). Despite its synthetic value, this reaction is considerably overlooked and underdeveloped. This is partly because of the difficulty of controlling the addition of a nitronate to a nitroalkene. The initial product formed is the nitronate anion 68, which is sufficiently reactive to also undergo conjugate addition to the nitroolefine 67, thus resulting in a mixture of oligomerization products48 including 70 and its higher congeners (Scheme 23), and in significantly reduced yields of the desired 1,3-dinitro products 69. Perhaps because of these difficulties, very few protocols for the above reaction have been published and are limited to one unasymmetric and two asymmetric methods 75 only. Thus, the addition of nitroalkanes to nitroolefins is one of the least developed and most intriguing types of Michael addition processes. Scheme 23. The conjugate addition of nitroalkanes to nitroolefines as a useful route for the construction of valuable functionalities R'V N02 R N02 L H+ R N02 H2 R NH? —* . .. - L ----- - L + FAA/N02 R e 2 R, N02 WC n' NH2 67 53 69 71 ‘R&NO2 E X i x i R N02 R O R NH2 R' N02 . OH , OH N02 R R 70 5.2 PROTOCOLS FOR THE CONJUGATE ADDITION OF NITROALKANES TO NITROOLEFINS Protocols for the non-asymmetric conjugate addition of nitroalkanes to nitroolefins are extremely rare, although some early reports on the Michael addition of nitroalkanes — or their alkali metal salts — to nitroolefines (either isolated or generated in situ) had been described“9 in the late 40’s to early 50’s. The conditions for these reactions were harsh and the yields were variable and often low. According to the best of our knowledge, the only representative and relatively efficient method reported so far is reported by Alcantara’s group in 1996”. According to that paper, reactions with the very reactive B-nitro esters as nucleophiles are fast in hexane or benzene, but less reactive aliphatic nitroalkanes are more sluggish, giving large amounts of polymers. Good yields of the desired 1,3- dinitroalkane products from these nitroalkanes can be obtained in reasonable reaction 76 times only if a large excess of nitroalkane (~110 eq) is used in a neat reaction (Table 22). The corresponding reactions in hexane or benzene are very slow, and in most cases the reactions are not diastereoselective, leading to mixtures of isomers. Table 22. Alcantara’s method for the synthesis of 1,3-dinitroalkanes R3 N02 R1/\I/ N02 + R3 V N02 Eth L/ N02 92 ~1 1O eq. rt R1 R2 entry Rl R2 R3 time(h) solvent yield% 1 Gal H H 20 none 72 2 Gal H Me 24 none 68 3 Ph H H l .5 none 5 l 4 Ph H Me 7 none 95 5 Ph H COOMe 2 hexane 66 6 Ph Me H 26 none 98 7 4-MeO-C6H4 Me Me 24 none 80 8 furyl Me Me 120 none 72 9 isopropyl H COOMe 3 hexane 79 The breakthrough came in 2006 when Du’s group reported a highly enantioselective version of the reaction using Cz-symmetric tridentate bis(oxazoline) and bis(thiazoline) zinc complexes“. This group had used these catalysts before in the context of Lewis acid catalyzed Henry reactions of nitromethane with a-keto esters“. The ability of these zinc species to activate nitroalkanes was applied successfully in the conjugate addition of the latter species to a series of substituted nitrostyrenes, affording the first 77 enantioselective protocol for the reaction with up to 95% ee for the resulting 1,3-dinitro compounds (Table 23). Table 23. The first asymmetric addition of nitroalkanes to nitroolefines 8?. 9 N H O N NL/JO S N N\_/_/S \_j\Ph Pfi' Lkph Ph“ 72 73 10 mol% 72 R “~02 + RZYNOZ 25 mol% EtQZn ; R2 A. 3N02 R1 R 80 mol% Ti(OiPr)4 LNG 3 toluene/hexane. R1 2 4 eq. 3 days, rt entry R1 R2 R3 yield% d.r. %ee 1* Ph Me H 87 1 1.7/1 95 2 4-MeC6H4 Me H 83 6.1/1 91 3 4-MeOC6H4 Me H 83 6.5/ l 89 4 2-MeOC6H4 Me H 90 5.8/1 91 5 3 ,4-(MeO)2C6H4 Me H 67 9 .6/ l 86 6* 4-FC6H4 Me H 79 9.3/ 1 92 7 2-FC6H4 Me H 87 6.2 90 8 2-C1C6H4 Me H 83 3 .8/ l 91 9 2-naphthyl Me H 83 3 .8/ l 88 10 2-furyl Me H 67 6.7/1 88 l 1 PhCHzCH2 Me H 80 3 .4/ l 72 12 Ph Et H 88 4.1/1 88 13 Ph Me Me O - - *Ligand 73 was used. 78 As can be seen from the data in Table 23, Du’s method is highly efficient for the synthesis of enantioenriched syn-1,3-dinitro products. The reaction is broad in scope since either electron rich or electron deficient nitrostyrenes can be used, and additionally, heteroaromatics (entry 10) and aliphatics (entry 11) can be also tolerated. Branched nitroalkanes though turned out to be sluggish nucleophiles (entry 13). Figure 17. Proposed mechanism for Du’s asymmetric addition N Ph 3 ,. S 0' \ 1" \\/Et/ 0 A/Nizn“.zn‘~6N\2 \ ' O , s 0‘ ’0 \“ Ph N k a ‘2 R/ R1 The conjugate addition of nitroalkanes to nitroolefines is proposed to work through the model depicted in Figure 17 (although the authors ignored the role of Ti), with the Zn activating nitroethane for deprotonation and at the same time forming a complex with the nitroolefin, which is oriented in the chiral pocket so that its Re face is exposed for nucleophilic attack“. The second protocol for the direct conjugate addition of nitroalkanes to nitroolefines was reported by Wang’s group in the same year”. It described the addition of branched nitroalkanes — mainly 2-nitropropane — to nitrostyrenes under neat conditions using the modified C inchona alkaloid 74 to obtain 1,3-dinitro products in good yields and modest to good enantioselectivities (67—85%ee, Table 24). This particular method has a major drawback: the catalyst is not reactive enough to provide reasonable reaction times even without the use of a solvent and under a large excess of the nitroalkane. Several days 79 are required for the reactions to go to completion (presumably the branched nitroalkane is too hindered to be a reactive nucleophile). Table 24. Wang’s protocol using catalyst 74 /\/ N02 + \r N02 10 mol% 74 £02 Ar ——’ neat, 0°C Ar N02 56 eq. Ar t(days) %yield ee% Ar t(days) %yield ee% Ph 6 79 78 2,6-C12C6H3 7 75 80 4-FC6H4 7 76 73 4-MeOC6H4 12 78 67 4-C1C6H4 8 70 83 2-MeOC6H4 8 73 84 2-CF3C6H4 9 78 83 2-BnOC.,H4 7 80 77 2-C1C6H4 10 73 8 1 2-thiophene 10 78 7O Finally, Maruoka’s group has published the enantioselective addition of preformed silyl nitronates to nitrostyrenes, a surrogate to the reaction of interest that requires the separate preparation of the highly reactive silylated nitroalkanes for use as nucleophiles“. Using catalyst 75, syn products can be obtained in excellent yields and good to excellent enantioselectivities. As may be expected for the reactions of these highly reactive species, very low temperatures (-78°C) have to be used in order to achive the maximum inductions. 80 Figure 18. Maruoka’s catalyst for the formal addition to nitroalkenes F3C I CFS Table 25. Maruoka’s formal addition using catalyst 75 r 93"“93 75 (0.5-2 mol%) H30+ EKG/02 Rz/VNOZ + F? [1L6 THF,-78°C,1h R2“ N02 ‘2... entry Rl R2 mol.% 75 yield% d.r. %ee 1 Et 4-MeOC.,H4 l 99 >95/5 91 2 >> 2-MeOC6H4 0.5 99 >> 87 3 >> 2-ClC6H4 0.5 98 >> 80 4 >> 2-naphthyl 0.5 95 >> 89 5 >> furyl 2 92 >> 93 6 >> thiophenyl 2 85 >> 91 7 >> cyclohexyl 1 99 >> 92 8 >> n-hexyl 1 93 >> 76 9 Pr Ph 1 99 >> 91 10 MeOCH2 Ph 1 99 82/18 83 81 5.3. THE CONJUGATE ADDITION OF NOTROALKANES TO NITROOLEFINES BY MEANS OF THIOUREA CATALYSIS It is clear from discussion in the previous section that the addition of nitroalkanes to nitroalkenes is a reaction that is just starting to be developed, and that there is plenty of room for improvement. There are a lot of aspects to address and improve, namely the problem of oligomerization, the diastereoselectivity in the non-asymmetric reaction, and the enantioselectivities, that are modest in the case of Wang’s organocatalyzed protocol”. The development of an efficient protocol for a diastereoselective non-asymmetric synthesis of 1,3-dinitro compounds, as well as the identification of catalyst structures — either metal based or metal free - that can induce high asymmetric inductions in these products would be highly desired. Our interest in the addition of nitroalkanes to nitroalkenes — as well as in the field of thiourea catalysis — focused our approach towards the development of an efficient protocol for this reaction in an investigation of the use of thioureas as catalysts. It is known that thioureas can activate nitrostyrene (and other nitroolefins) for the addition of a vatiety of nucleophiles". On the other hand, the activation of nitroalkanes and the enhancement of their acidity through their interaction with thioureas is also known, in several reports where the successful employment of nitroalkanes as reactive nucleophiles is described”. Despite all these reports, a thiourea promoted addition of nitroalkanes to nitroolefines has never been reported in the literature, although it would be expected that the application of thioureas could facilitate the above transformation and possibly control the selectivity of the addition (Scheme 24). A careful review of the field of thiourea catalysis reveals that, concerning the employment of thioureas as catalysts, this particular 82 transformation is probably one of the few reactions involving nitro compounds that has not been addressed so f ‘3‘. Scheme 24. The conjugate addition of nitroalkanes to nitroolefines promoted by thioureas S 3 ft Ft R R R FL ,R ‘ ’ ‘ ' N N ANA H ”A H AA AA N02 N02 _= : ——* _= _: _= : —* R' \+, \+, O,+,,O Fl" 0 ND 0 N O N .cH ./” WI I=’\ Our approach is depicted in Scheme 24. It is expected that the use of thioureas for the above reaction will activate the nitroalkane, enhancing its acidity, and activate the nitroolefine, enhancing its reactivity of the C-C double bond. The presence of a catalytic amount of a thiourea - for substrate activation - and the presence of a base - for nitroalkane deprotonation — should be sufficient for reaction acceleration, and might lead to a chemoselective and diastereoselective reaction outcome. The chemoselectivity is an important issue since polymerization can cause a decrease in yield of the Michael adduct or require large amounts of nitroalkane to maintain good yields. The rest of the chapter describes the use of thioureas in the context of the conjugate addition, in an investigation that was focused on i) the employment of simple aromatic thioureas for the promotion of the unasymmetric reaction and ii) the development of a novel chiral thiourea catalyst for a highly enantioselective conjugate addition methodology. 83 5.3.1. Development of the thiourea catalyzed non-asymmetric addition It was realized that an investigation of the thiourea promoted addition of nitroalkanes to nitroolefines should be first focused on the non-asymmetric reaction, and on the effects that simple achiral thioureas would have on the transformation, since as it was mentioned, the employment of thioureas in this particular type of reaction has never been described before in the literature. The small symmetrically substituted aromatic thioureas 76-78 shown in Figure 19 have been reported before in the literature in various studies conducted by Schreiner’s group and other research groups, for the acceleration of several transformations including Diels-Alder, Bayllis-Hillman, or Claisen rearrangements“. Schreiner has established that diaryl thioureas of this type with electron withdrawing substituents on the meta and/or para positions are the most effective candidates for rate acceleration in a diverse set of reactions343". Thus, it was judged that this type of thioureas should be the focus of our study. Additionally, we decided to also make and evaluate the bis-thiourea 79, based on the idea that two thioureas incorporated in the same structure might result in activation of both reactants. Figure 19. The substituted aromatic thioureas used in this work F30 NYN cr=3 cr=3 “A,“ O s O S O Q s Q JL F 77 F N N CF3 Cl H n Cl CF 1“ . 7! 7i? Q r O or 78 cu Scheme 25. Synthesis of aromatic thioureas 47, 76-79 thiophosgene/ H H o EI3N N N 47, 40 /o —» A , \A 76, 60% ArNH2 r \[r r 77’ 40% THF 3 78, 96% CF3 CENHz scn’ : ‘CF3 NH2 THF 1 00% 79 The synthesis of thioureas 47 and 76 to 79 was accomplished using standard literature procedures .4" As shown in Scheme 25, thioureas 47, 76-78 can be accessed from the reaction of commercially available substituted anilines with thiophosgene and triethylamine in THF. On the other hand, thiourea 79 can be made from 1,2- diaminobenzene with 3,5-bis-trifluoromethylphenyl isothiocyanate. We started our investigation by looking at the effect that the thioureas 47 and 76- 79 have on the rate acceleration of the conjugate addition. The reaction between trans-B- nitrostyrene and l-nitropropane was used as a model reaction in CDCl3, employing only 5 equivalents of nitropropane, a very small excess compared to Alcantara’s protocol, and using 10 mol% thiourea and 10 mol% triethylamine. The rate of the reactions was followed by lH-NMR using the catalyst as an internal standard. Thiourea 78 was not soluble in CDCl3 and therefore was not evaluated. The results for reaction with thioureas 47 , 76 and 77 are shown in Scheme 26. As can be seen in Scheme 26, the reaction without thiourea (using 10 mol% Et3N) is very slow, affording only a 5% yield of the desired 1,3- dinitroalkane after 3h. On the other hand, the rate acceleration that thioureas 47, 76, and 85 77 can cause is very significant: Thioureas 76 and 77 can increase the relative rate of the reaction by approximately 5 and 6 times, respectively. It is obvious that thiourea 47, bearing the 3,5-bis-(trifluoromethyl) moiety is the most efficient catalyst, leading to almost quantitative conversion of the starting nitrostyrene in 2 hours, and affording ~60% yield of the desired 1,3-dinitroalkane. Scheme 26. Rate accelaration of the conjugate addition using thioureas 47, 76, 77 \ N02 10 mol% cat. N02 10 mol°/oEth N02 + ——> /\/N02 cool3 5 eq. yield (as) 0 20 40 60 80 100 120 tlme (mln) The results shown in Scheme 26 were very promising at that point, because it showed the potency of thioureas in promoting the coupling between the two nitro- components. The diastereoselectivity of the reaction was not determined — the yields in Scheme 26 refer to that for the combined syn and anti products. Since good yields can be 86 obtained in only two hours these results prompted further investigation of these thioureas, especially 47. Next, the efficiency of thioureas 47 and 76-79 was investigated in regard to the yield and the diastereoselectivity of the addition of various nitroalkanes. The results for the reactions of nitroethane, l-nitropropane and l-nitrobutane with nitrostyrene are shown in Table 26. Nitroethane affords a mixture of diastereomers (entry 1), however it was delightful to observe that the control of the selectivity for the rest nitroalkanes is much better. All thioureas can control the selectivity of the reaction in favor of the syn isomer, furnishing the desired product in good yields (up to 83%) and affording d.r. ratios between 77/23 and 87/13. The background reaction without thiourea (10 mol% Et3N, entry 7) is very slow, leading to a trace of the product after an overnight period. It was interesting to observe that the background reaction even with a stoichiometric amount of base is still inefficient in promoting the reaction and only furnishes a 39% yield with a slight preference for the anti isomer, revealing that the catalytic and the background reaction complement each other in terms of stereocontrol (Table 26, entry 8). Finally, the reaction with nitrobutane (entry 9) is consistent with the rest, giving 82% yield of the syn-enriched product. 87 Table 26. The conjugate addition of nitroalkanes to nitrostyrene catalyzed by thioureas (a) \ N02 R N02 catalyst ©/\/ EtSN N02 # + Fl NO2 Toluene v r.t. 80a yield% d.r. (syn/anti) entry R catalyst (mol%) mol% Et3N (b) (C) 1 Me 47 (10) 10 85 55/45 2 Et 47 (10) 10 83 77/23 3 Et 76 (10) 10 66 78/22 4 Et 77 (10) 10 58 87/13 5 Et 78 (10) 10 75 80/20 6 Et 79 (10) 10 67 84/16 7 Et - 10 7 n.d. 8 Et - 100 39 40/60 9 n-Pr 47 (10) 10 82 77/23 (a) Reaction conducted with 0.22M with respect to nitrostyrene in toluene for 12h using 10 equivalents of nitroalkane. (b) Isolated yield of the mixture of diastereomers. (c) Determined by HPLC analysis. It was interesting to observe that in all of the reactions in Table 26 all of the thioureas can promote the formation of the desired product and at the same time suppress the production of polymeric material. The excess of nitroalkane used in the above reactions was only 10 equivalents, much lower than in Alcantara’s protocol (~l 10 eq.).50 Nonetheless, the rate acceleration that thioureas afford in favor of the desired product is 88 apparently high enough to provide quite clean reactions (in terms of the amount of polymer formed). The reaction catalyzed only by Et3N (entry 8) gave a higher amount of polymeric material, as judged by the messy crude lH-NMR spectrum obtained. Since catalyst 47 turned out to be the optimum for the reaction, the addition of l- nitropropane to nitrostyrene catalyzed by thiourea 47 was next examined in different solvents and using various catalyst loadings in order to optimize the conditions and the results are summarized in Table 27. The yields of the products were determined by the 1H- NMR spectrum of the crude reaction mixture with triphenylmethane as internal standard. The catalyst loadings for the reactions in toluene were gradually reduced as indicated in entries 1 to 4. The diastereoselectivity of the reaction remains intact when the catalyst is reduced from 10 to to 2 mol%, while a small drop-off is observed in the rate of the reaction (entry 3, 58%). This can be suppressed by adding more nitropropane and increasing the reaction time. For example, by using 20 equivalents of nitroalkane and keeping the catalyst loading at 2 mol%, a very good yield of the product can be obtained (85%, entry 4) in 2 days reaction time and no decrease is observed in the syn to anti ratio. The reaction in other solvents also afforded good yields and d.r.’s, with THF being the only exeption, which is as expected for a solvent that can function as an H—bond acceptor and decelerate the reaction (entry 8, 45%). 89 Table 27. Optimization of conditions for the reaction using thiourea 47 (a) NO \ N02 thiourea 47 2 ————> N02 Et3N + /\/ N02 solvent, rt 80a entry solvent mol% 47 mol% Et3N eq. n-prNO2 yield% (b) d.r. (c) l toluene 10 10 10 83 77/23 2 toluene 4 4 10 63 80/20 3 toluene 2 2 10 58 ((1) 79/21 4 toluene 2 2 2O 85 77/23 5 - 2 2 20 82 73/27 6 benzene 10 10 10 89 68/32 7 DCM 10 10 10 78 75/25 8 THF 10 10 10 45 77/23 (a) Reactions conducted in the corresponding solvent, for l2-15h, except entries 2-5 which were run for 2 days. Entries 1-5 were run with 0.22M with respect to nitrostyrene, entries 6-8 were run with 0.1M with respect to nitrostyrene. (b) Determined by lH-NMR of the crude mixture. (c) Determined by HPLC analysis. ((1) Incomplete conversion. After optimization of the solvent and catalyst loading, the conditions on Table 27, entry 4 were chosen as the optimum and the scope of the reaction was examined for a series of differently substituted nitrostyrenes. As can be seen from the data in 90 Table 28, a number of differently substituted nitrostyrenes can undergo the thiourea catalde addition to give the syn 1,3-dinitroadducts 80a-i in good yields. Electron rich nitrostyrenes (entries 2 to 4) were less reactive and needed longer reaction times (40h) to go to completion, while electron deficient substrates (entries 5 to 9) afforded complete conversions in only 16 hours. In every case, the starting material underwent complete conversion (90-100%) with only small amounts of polymer formed, and the reactions were in general very clean compared to the background reaction (Table 26, entry 8). It seems that the thiourea catalyst is accelerating the formation of the desired adducts suppressing the overaddition process and affording very good turnovers. Additionally, the reactions favor the formation of the syn diastereomer in good selectivities that range from 3/1 to 9/1. 91 Table 28. Reaction scope for the thiourea catalyzed addition of l-nitropropane to [3- nitrostyrene (a) 2 mol% H H F30. ; ,N\n/N. ; ,cr-‘3 S CF CF 23 I E N 3 N02 R/V N02 + /\/N02 mo /° t3 : lNOz (O'ZZM) 20 eq toluene, rt BOa-I yield% syn entry R adduct overall yield% d.r. (c) (b) 1 Ph 80a 87 75 77/23 2 4-MeO-C6H4 80b 82 73 75/25 3 2-MeO-C6H4 80c 83 66 80/20 4 4-Me-C6H4 80d 72 56 77/23 5 2-Cl-C6H4 80c 80 68 85/ 15 6 4-Cl-C6H4 80f 73 58 80/20 7 4-Br-C6H4 80g 66 57 86/ 14 8 3-Br-C6H4 80h 75 69 92/8 9 2-Br-C6H4 80i 66 6O 90/ 10 (a) Reactions run overnight except entries 2-4 which were run for 40h. (b) Isolated yield after column chromatography. (c) Determined from the ratio of isolated products except entries 1 and 2, which were determined by the 'H-NMR of the crude reaction mixture. An explanation for the preferential formation of the syn isomer can be obtained from the examination of the corresponding transition states for the two diastereomeric dinitroalkanes (Scheme 27). It seems that the lowest energy transition states for both the syn and the anti product formation would be the ones where the groups with the largest A 92 values would be anti to each other. That means that the phenyl group of the nitrostyrene (A=3.l kcal/mol) and the ethyl group of the nitropropane (A=l .9 kcal/mol) have to adopt an anti relationship (the A value of the NO2 group is 1.1 kcal/mol). In the TS A, which gives the syn product, this requirement also results in the nitro groups being opposite to each other. On the contrary, considering the TS B that gives the anti product, the anti relationship of the phenyl and the ethyl group results in the nitro groups coming close to each other and pointing to the same direction, creating a steric hindrance between the two thioureas. The steric congestion of the TS B seems to be a plausible explanation for the observed stereoselectivity, although more experiments should be done to verify this hypothesis (e.g. rate studies to prove that the reaction is second order in catalyst). 93 Scheme 27 . A possible explanation for the preference of the syn diastereomer formation in the reaction catalyzed by thiourea 47 F3C©CF3 N -, F30 8:2 9 ,CH3 F3C H‘O’NQH CF3 NO ”IO-HM): —’ N20 CF. m ‘O-HN S 2 CF TSA (favored) F C 3 3 syn F30 CFS H c “N S CF3 3 (5 \_/ \ HN .‘ '_ ’ A \NO HYN O F C CFS 2 N—0 —> No2 at F30 3 anti TS B (unfavored) 94 5.3.2 Development of the asymmetric conjugate addition of nitroalkanes to nitroolefines The success of thiourea catalyst 47 in promoting the syn-selective non-asymmetric addition of l-nitropropane to fi-nitrostyrene prompted the investigation of an enantioselective variant using chiral thiourea catalysts. The reaction design for the asymmetric variant is shown in Schemes 28 and 29 and includes two different approaches. a) Bis-thiourea catalysts that can mutually activate both reaction components (Scheme 28). In this approach, one thiourea moiety is used for H-bond activation of nitrostyrene and the second thiourea moiety is activating the nitroalkane for deprotonation. In this case an external base will have to be used for the deprotonation, leading to the formation of a nitronate, in a complex that will feature the two components in close proximity for reaction. This concept was the one that was used for the development of the asymmetric aza-Henry reaction in Chapter 4. Scheme 28. Bis-thiourea approach for the conjugate addition JSL S N N’Ar NJL N Ar / H H K H H 8.. a b g,- chiral ‘ N ’ B - hiral ‘ N ’ . . deprotonation c scaffold R’kfi' / _ scaffold RJ \ JSL \ S N N,Ar NJLN Ar '7 ti H H 5 i a a .+, , ‘ O‘N O O‘N'O b) A bifunctional thiourea catalyst that will incorporate the base in its structural framework (Scheme 29). In this approach, that has been used by Takemoto and other groups for other transformations”, the nitroalkane will be deprotonated by the basic moiety of the catalyst probably through H—bonding activation from the thiourea. The nitroolefin could then displace the nitronate and interact with the thiourea, while the nitronate anion should be held by the positively charged basic moiety through H-bonding, in a complex that would again feature the two reaction components in close proximity for reaction. The investigation to be described in this chapter was thus focused on identifying efficient catalysts for both approaches. Scheme 29. Bifunctional thiourea approach for the conjugate addition it . N N’Ar N’lLN Ar / H H K H H chiral Q‘fi '0 deprotonation chiral O, N ,0 ‘ scaffold /L\—H ' scaffold I \ ‘ “ r .B. ) \BH 4. IRrA/NOZ it N N'Ar K H H R N02 chiral 0°N’O LNG ‘ _ scaffold RA Ar 2 \ Fl' BH _ O + O‘Nl M 96 5.3.2.1 Identification of the optimum chiral thiourea catalyst Concerning the first approach, involving bis-thiourea catalysts, the bis-thioureas 52 and 55 were evaluated for the reaction of B-nitrostyrene with l-nitropropane and 1- nitroethane. These reactions were studied in various conditions and the results are shown in Table 29. Table 29. Reactions of nitroalkanes with B-nitrostyrene using catalysts 52 and 55 \ N02 N02 W 20 mol% cat. N02 + RVN02 solvent, temp. (10 eq.) entry R catalyst (%) Et3N (%) solvent T(°C) yield% d.r. (syn/anti) %ee 1 Et 52 (20%) 20 DCM 25 80 70/30 23 2 Et 52 (20%) 20 DCM -20 80 85/ 15 22 3 Et 52 (20%) 2O toluene -80 25 81/19 34 4 Et - 40 toluene 25 35 56/44 - 5 Et 52 (20%) 5 toluene 25 80 90/10 19 6 Et 55 (20%) 5 toluene 25 100 80/20 17 7 Et 52 (20%) 0 toluene 25 0 - - 8 Me 52 (20%) 20 DCM -20 100 63/37 0 97 As can be seen in from the data in Table 29, the reaction with catalyst 52 at rt afforded a promising yield for the syn isomer but the enantioselectivity was disappointing (23%, enatry 1). An effort was made to increase the induction by lowering the temperature to -80°C (entries 2, 3) but with no result other than a dramatic drop-off in the rate of the reaction. The background reaction was then measured to determine whether it is responsible for the low inductions, and it was found that 40 mol% Et3N can afford ~35% product at rt. The background reaction was slower than the catalyzed reaction, but nevertheless it was significant. In order to diminish its effect, lower amounts of EtsN were tried (5 mol% with catalysts 52 and 55) but these efforts did not improve the selectivity. In the absence of Et3N no reaction was observed, which means that the catalyst itself is not able to promote the deprotonation of the nitroalkane to the nitronate. Finally, the reaction with nitroethane was unsuccessful as well, affording racemic product. From the above results it became evident that the first approach that was to utilize bis-thioureas as catalysts was not successful, since neither 52 nor 55 was able to afford satisfactory inductions for the Michael addition product. The investigation then switched gears to the second approach, that is to the identification of efficient bifunctional catalysts. 98 A careful look into Wang’s development of an organocatalytic protocol for the Michael addition of nitroalkanes to nitroolefins52 revealed that bifunctional thioureas could actually be promising candidates for the promotion of the enantioselective conjugate addition. A part of his screening of catalysts included the bifunctional catalysts 66 and 81 in a neat reaction without solvent. As can be seen in Table 30, catalyst 66 (which is a catalyst Wang’s group has developed for other asymmetric reactions) affords a modest but promising enantioselectivity but the reactivity of the catalyst is not sufficient. On the other hand, catalyst 81 is obviously more reactive,but the increased reactivity does not translate to increased selectivity. Table 30. Wang’s catalyst screening H M80 S CF3 CD . N02 10% t. \ N02 + YNOZ ca N02 56 eq. neat, rt catalyst yield% %ee 66 29 48 81 73 30 The above results that Wang’s group reported are interesting because they reveal that a thiourea catalyst could be efficient for the addition of nitroalkanes to nitroolefines, if it incorporated a base that would be strong enough to promote the nitroalkane 99 deprotonation. It is obvious that 81 is reactive enough and this may be attributed to the strong basicity of its nitrogen. Wang's catalyst, on the other hand, suffers from the low basicity of the nitrogen that is in conjugation with the binaphthyl ring. The above considerations inspired an interesting idea: it was envisioned that the reactivity problem associated with the low basicity in Wang’s catalyst (66) could be solved if the catalyst was derivatized with a more reactive base. Along these lines, catalyst 82 was designed, that features a dimethylaminopyridine (DMAP) moiety in its structure. Catalyst 82 should be more reactive than 66 simply on the basis of the higher basicity of the DMAP moiety compared to that of the dimethylaniline in 66 (Figure 20). Figure 20. Design of a novel bifunctional catalyst based on BINAM Wang's catalyst: DMAP analog: Low basrcrty higher basicity zit/£1310 OHAHO """ \NfipKa-w -12 p-Ka:‘52. 2 (H20)O) _ + / \ / R ,0" ’ ' HN _ N\ Aminopyridine moiety: \=N\ deprotonation of nitroalkane/ + o - - HN\ interaction via double H-bond ' H“ In addition to the greater basicity of catalyst 82, another unique feature of this catalyst is that after deprotonation of nitroalkane the 2-aminopyridinium moiety would be 100 expected to interact with the resulting nitronate through double H-bonding, possibly resulting in a well-defined orientation in the chiral pocket. The ability of 2- aminopyridinium ions to interact with nitro groups is documented in recent literature reports. For example, Takenaka’s group found in 2007 that 2-aminopyridinium ion 83 can activate nitroalkenes through double H-bonding toward the addition of N-methylindole and the Diels-Alder reaction with cyclopentadiene”. Scheme 30. 2-Aminopyridinium catalyst for nitroalkene activation 83 (dos-0.5%) N R“ N02 > )\/ N02 — Ar-H R TFPB 990/ / / \ up to o ‘N 83 (ODS-0.5%) R 83 liq/V N02 : / Q N02 up 10 54°/o endo/exo: up to 30/1 In the case of catalyst 82, the aminopyridinium ion of the catalyst could interact via H-bonding either with the nitroolefin or the nitronate ion, but it is anticipated that it will most likely form an ion pair with the latter, which is negatively charged and thus more electron rich. Catalyst 82 can be synthesized trough a simple two-step sequence starting from the commercially available (R)-BINAM 51. The first step involves a Buchwald amination of 2-chloro-4-dimethylaminopyridine using Pd catalysis. The reaction using a 1:1 stoichiometry between (R)-BINAM and 2-chloro-DMAP 85 gives 60% isolated yield of the desired monosubstituted product 84, along with a small amount of disubstituted adduct 101 that can be easily separated with column chromatography. 2-Chloro DMAP can be obtained from 2,4-dichloropyridine through an a nucleophilic aromatic substitution using dimethylamine. Scheme 31. Synthesis of catalyst 82 F C NCS CF szdbaa 3 3 :23 00 ‘ NaOtBu H ”“2 35(1 90) CF3 “JLN CFa Owl-12 ..N N Toluene DCM, rt ' H l \ 82:00 60°/o / \N/ A \ 0 5° C Cl N/ 820/0 85 The second step involves the formation of the thiourea moiety from the reaction of 84 with 3,5-bis(trifluoromethyl)phenyl isothoocyanate. The product 82 can be obtained in 60% isolated yield after chromatography purification. The preliminary results from the evaluation of the new catalyst were very promising. In terms of reactivity, the data in Table 31 present a comparison with catalyst 66: catalyst 66 showed no reactivity for the given set of conditions (10 equivalents nitroalkane in dichloromethane at r.t.) while 82 provided the dinitro adduct 80a in an excellent 90% yield as determined by NMR spectroscopy. Note that in the absence of solvent catalyst 66 can promote the addition of a nitroalkane to a nitroalkene to a small degree. A preference for formation of the syn diastereomer was observed along with 78% ee, which at this point 102 was very satisfactory. Reactions in other solvents revealed that benzene is actually the optimum in terms of the stereoinduction (80%ee, entry 4). Table 31. Evaluation of the new catalyst 82 in the conjugate addition (a) \ N02 NC)2 ©/\/ 20 mol% cat. N 0 g 2 solvent, temp. + /\/N02 (10 9%) 80a entry catalyst solvent yield% (b) 80a d.r. (c) (syn/anti) % ee (c) 1 66 DCM O - - 2 82 DCM 90 59/41 78 3 82 toluene 90 63/37 78 4 82 benzene 88 64/36 80 (a) Reactions run with 0.1M with respect to nitrostyrene for 12h. (b) Determined from the crude lH-NMR spectrum using triphenylmethane as internal standard. (c) Determined by chiral HPLC. After this study, catalyst 82 was chosen as the optimum catalyst for the reaction and an effort to improve the result through an investigation of various conditions was conducted. 5.3.2.2 Preliminary efforts for optimization The effort towards optimizing the conditions for the reaction of B-nitrostyrene with nitroalkanes using catalyst 82 begun with variations in the temperature of the reaction with l-nitropropane. Specifically, lowering the temperature to -25°C was attempted in order to achieve higher selectivities. Unfortunately this had a dramatic effect on the rate of the reaction which decreased significantly and resulted in only 31% yield (Table 32, entry 2). The opposite would be expected when warming up the reaction, however, in the case of 103 this particular transformation there is one complicating factor, the formation of the polynitrostyrene (discussed previously). It tumes out that increasing the temperature to 50°C causes an increase in the rate of the polymer formation which is apparently higher than the increase in the rate of the formation of the desired product. This seems to be the explanation for the result in entry 3. The reaction at 50°C actually gave complete conversion of the starting material, but with a much lower yield of the desired product that at room temperature. The decrease in the mass balance is assumed to be due to the formation of polymer and this is consistent with the complicated NMR spectrum that was observed for the crude reaction mixture. Unfortunately, temperature seemed to be a limiting factor and it was obvious that any further optimization had to be conducted only at room temperature . 104 Table 32. Preliminary optimization of conditions using catalyst 82 (a) m N02 R N02 20 mol. °/o 82 = N02 + Fl VN02 solvent, temperature (10 eq.) entry R solvent T(°C) yield%* d.r. %ee syn % ee anti 1 Et benzene rt 80 60/40 82 7O 2 Et toluene -25 31 n .d. n.d. n .d. 3 Et benzene 50 54 50/50 80 7O 4 Et neat rt 100 50/50 78 72 5 Et benzene, 4A M.S. rt 62 45/55 40 44 6 Me benzene rt 100 57/43 62 6O 7 Pr benzene rt 96 55/45 8 1 79 8 Et benzene rt 75 66/33 82 n.d. (a) Reactions run with 0.1M with respect to nitrostyrene for 12h. *Entries 1-7: lH—-NMR yield using triphenylmethane as internal standard. Entry 8: Isolated product after silica gel chromatography. Two more experiments were conducted with l-nitropropane: a neat reaction without any solvent (entry 4), in order to increase the rate and the yield, and a reaction using molecular sieves, in an effort to ensure dry conditions and removal of any traces of moisture that would interact with the catalyst and/or the nitrocompounds (entry 5). Unfortunately, neither change was helpful, since the neat reaction was too fast and unselective in terms of the dr. ratio (entry 4), and the molecular sieves had a diminishing 105 effect on the selectivities (entry 5). The latter was not expected, and the effect of the sieves on the reaction outcome is not clear. Finally, a few more nitroalkanes were tried in the reaction at r.t. Nitroethane (entry 6) gave a lower but still respectable enantioselectivity (62% for the syn product) along with excellent yield, while l-nitrobutane (entry 7) afforded results similar to those with l- nitropropane, showing that the reaction is not very substrate sensitive (at least concerning the nitroalkane) and can be applied to different nitroalkane nucleophiles. Among the last two substrates, l-nitropropane was chosen for further study because of the easier separation of the enantiomers in the chiral HPLC. It should be noted here that for all the experiments in Table 32 the yield that is reported is determined from the crude 'H-NMR spectrum using triphenylmethane as a standard. Specifically, a known amount of triphenylmethane was added to the reaction mixture at the end of every reaction and the yield was calculated by integration of aliphatic protons of the product and the methine proton of the standard that has a characteristic absorption at ~ 5.5 ppm. This accuracy of this method was checked by performing a couple runs where the product was isolated by chromatography. The data in entry 8 refers to isolated product and is the average of two runs. The consistency with the data in entry I allowed for the reliable use of triphenylmethane as a standard in further experimentation and optimization. In addition, the high polarity of the catalyst allowed for a convenient determination of the ee’s and d.r. ratios without the need of column purification: All that was needed in order to obtain this data was to simply filter the crude reaction mixture through a short Pasteur pippete containing a small amount of silica gel. The polymer and the catalyst stayed on the top of the silica, and product of very good 106 purity was eluted, which allowed for very clean HPLC spectra. This method proved to be very useful in rapidly optimizing the reaction without the need to perform column chromatography for each and every entry, thus saving a lot of time. The method described here was applied to all the experiments that are discussed in the following sections, except of course the section that deals with the substrate scope, where isolation and purification of products was required. A second point of importance is the reactivity and the chemoselectivity of the catalyst 82. For the reactions conducted at r.t. the conversion is always nearly quantitative, with only small amounts of polymer formed (<20%). The chiral thiourea seems to activate the starting materials in favor of the formation of the desired adduct suppressing the production of oligomers or polymers, just as the achiral thiourea 47 did for the non- asymmetric addition (Table 28). 5.3.2.3 Effect of catalyst loading: discovery and investigation of a highly unusual selectivity trend The experiments discussed in the previous section led to the conclusion that the optimum conditions for the maximum yield and selectivity of the reaction are met when the reaction is conducted at room temperature in benzene. Up to this point the experiments have been performed using 20 mol.% catalyst, along with 10 equivalents of l- nitripropane. The next objective thus was to try to reduce the catalyst loading in order to achieve a satisfactory balance between the amount of catalyst spent and the yields and selectivities obtained. The investigation was first focused on the effect of catalyst loading. A series of experiments were conducted using 10 equivalents of l-nitropropane in benzene at room 107 temperature. All conditions were kept consistent and a number of different catalyst loadings from 0.5% to 100% were screened. Analysis of the data led to a completely unexpected result from the above study: It was intriguing to observe that a reduction of the catalyst loading not only didn’t cause a drop in the selectivities, but on the contrary, it lead to an unexpected enhancement of the selectivities! As can be seen from the data in Table 33 a stoichiometric amount of catalyst affords a preference for the anti isomer, with ee’s in the 60’s and 70’s, while catalyst loadings from 20% and lower favor the syn isomer, with the asymmetric induction increasing very significantly. The enantioselectivity of the syn adduct spans a range from 70% (entry 1) to 95% where the catalyst is used in only 0.5 mol.% with respect to the nitroolefin. A significant increase in the diastereomeric ratio was also observed for reduced loadings, with an increase to 86/14 for the lowest loading (entry 7). This selectivity trend is obviously extremely unusual. Additionaly, the fact that catalyst loadings as low as 0.5% can induce selectivities as high as 95% demonstrates that thiourea 82 is a remarkably selective catalyst. 108 Table 33. An unusual selectivity trend observed for the catalyst loading study for 82 (a) CF3 S Q 03 min CF, .N N ©/\/ / NO2 (0.1M) /N\ mo2 r "' /\/ N02 Benzene,rt (10 eq.) 80° entry cat.load.% NMR yield% (b) d.r. (syn/anti) % ee (syn) % ee (anti) l 100 60 39/61 70 63 2 20 80 60/40 82 70 3 10 40* 71/29 84 60 4 5 70* 76/24 90 70 5 2 43* 82/18 94 n.d. 6 l <30* 87/ 13 94 n.d. 7 0.5 18* 86/14 95 n.d. (a) Reactions were run for 15h except entry 7 which was run for 40h. (b) Determined using triphenylmethane as internal standard. *Incomplete conversion It should be noted here that the rate of the reaction slowed down significantly when the catalyst was used in a loading of 10 mol% or less. For entries 3 to 7, the conversion was incomplete and a significant amount of starting material was unreacted. Nevertheless, it is worth mentioning that for those reactions with reduced loadings the crude 1H-NMR spectrum was generally cleaner with respect to polymer formation. In 109 general, the lower the catalyst loading that was used, the less polymer was formed and the cleaner the reaction mixture that was obtained. The trend that the enantioselectivity for the syn product followed with respect to the catalyst loading was very smooth, and this is apparent when plotting the induction for the product and the catalyst loading, as illustrated in Figure 21 . Figure 21. Plot of the asymmetric induction for syn - 80a versus the catalyst loading eo(syn)96 8 i 8 41 I on N 10 15 20 25 catkmdl% O 01 It is obvious that 0.5-2 mol% seems to be the optimum catalyst loading with respect to the selectivities, but unfortunately these amounts of catalyst are not practical because the yields of the product are not acceptable. The solution to this problem came with further optimization and will be described later on in the next section. 110 The investigation was then focused on rationalizing the above set of data in order to shed light on the mechanism of this transformation. A careful look into the selectivity trend shows that there are two factors that could possibly operate in the higher inductions observed upon decrease of the catalyst loading: (a) the decrease of the catalyst concentrations and / or (b) the increase of the nitropropane/catalyst ratios (since the catalyst amount is reduced across the table but the substrate amount is kept constant). Three different scenarions were envisioned in order to account for the unusual effects that these two factors could have on the selectivities: A) stereochemichal “scrambling” could happen if a nucleophile bound to one catalyst moiety reacts with an electrophile bound to another catalyst when the catalyst concentration (or loading) is high enough and catalyst moieties have a higher probability of coming close to each other in solution. This hypothesis is illustrated in Figure 22 as pathway A. If it is presumed that pathway D, which represents the design in the original model, is the more selective, then the scrambling of stereochemistry at high catalyst concentrations (pathway A) would occur because the nitroalkane would attack nitrostyrene in a stereochemically less defined way, or even attack the opposite side of the C-C double bond rather than the one that is exposed in D. According to this scenario, reducing the catalyst loading should disfavor pathway A and result in higher enantioselectivity. lll Figure 22. Possible scenarios to account for the catalyst loading effects A (Cross condensation) l lower cat. oonoentratlon F 3C CF3 Sj’NH .- O ,‘ / NH___O,N\/\Ar """NH _ _ _ Ot—N Jim” \ \N \ (+ / D (moraselecfivespecies) lower cat. ooncentratlon/ higher RN02lcatalyst ratio higher HNOg/cat. ratio CF3 C Mi? F300 HQ’ HNYS S I N . a ,‘tN’ ,x’HN O O b O _ —N\ B I 112 B) Intermolecular H-bond between different catalyst moieties that would lead to catalyst self-aggregation (species B in Figure 22) when the catalyst loading is high and the nitroalkane / catalyst ratio is small. This scenario seems reasonable on the basis of the potential of the electron-rich sp2 nitrogen of the DMAP moiety of the catalyst to accept an H-bond. As the nitroalkane/catalyst ratio increases (low cat. loadings) it is expected that the interaction of the catalyst with the substrates will be more favorable, and this interaction can break-up the aggregates and lead to species D. The magnitude of the nitroalkane/catalyst ratio indeed ecomes very large when switching from 100 mol% catalyst loading (ratio=10/1) to 20 mol% catalyst (ratio=50/1), to 5 mol% catalyst (200/1) and then to 0.5 mol% catalyst (2000/1), and so at these very high ratios it is expected that the catalyst would increasingly be more likely to associate with the substrates than with itself. The observed trend on the selectivities then can be explained if the aggregated species B can also catalyze the reaction but in a less enantioselective fashion than D. C) Intramolecular H-bonding of the catalyst as indicated in species C would leave only one thiourea N-H bond available for substrate activation, and this should presumably lead to less defined substrate orientations and thus to lower asymmetric inductions. Species C involves an intramolecular interaction and thus does not depend on the concentration of the catalyst itself, but its formation should be clearly dependent on the nitroalkane/catalyst ratio. Again, low catalyst loadings would translate to high nitroalkane/catalyst ratios, which should promote the break-up of the intramolecular H-bond (just as in the case of B) and the formation of the more selective species D. As with species B, the explanation of the observed trend would require that species C be able to catalyze the reaction with lower inductions than species D. 113 We next embarked on a series of studies in order to distinguish between scenarios A, B and C, and to probe the mechanism of this highly enantioselective addition. First, it is obvious that pathway A should be clearly concentration-dependent and that dilution of the reaction mixture should disfavor the formation of A and lead to higher inductions. A series of reactions were conducted where the same set of teagents and catalyst was dissolved in different amounts of benzene so that the catalyst loading and the nitroalkane/catalyst ratio were kept the same while the concentration of all the species including the catalyst were varied. The results of these experiments are depicted in Table 34. Table 34. Concentration study of the reaction (a) 20 mol% CF3 00 are. ..N N\ mmz 00 HEN? N02 k No2 + /\/N02 Benzene,rt 7 (10 eq.) 80a entry [catalyst], M yield%(NMR) (b) d.r. %ee (syn) %ee (anti) 1 0.040 77 58/42 84 70 2 0.010 71 64/36 85 70 3 0.005 55 52/48 78 64 4 0.003 45 55/45 67 48 5 0.002 31 55/45 70 42 (a) Reactions run for 48h. (b) Determined from the IH-NMR of the crude reaction mixture using triphenylmethane as internal standard. All entries gave incomplete conversion of nitrostyrene. ll4 The concentration study turned out to be very informative since the results were the exact opposite of what would be expected if pathway A was operating. When the reaction was performed at more diluted concentrations a very significant drop in the ee’s of both the syn and the anti products was observed. This clearly rules out pathway A. The dependence of pathway B on the concentration in not very clear at this point. An increase of the concentration may promote the formation of the aggregated species B, but on the other hand the increase in the concentration may allow for a more efficient interaction of the catalyst with nitropropane, thus disfavoring the aggregation. Thus, these two effects would offset each other. The concentration study also led to an important observation: An increase of the catalyst concentration is accompanied by higher inductions. This was not the case in the catalyst loading study (Table 33) where increased loadings — and thus, catalyst concentrations - gave lower inductions. This observation was the key in realizing that for the catalyst loading study, the factor operating in the catalyst loading trend is probably not the catalyst concentration itself, but rather, the nitroalkane/catalyst ratio. In other words, the nitroalkane/catalyst ratio was responsible for the unusual selectivity trend in the data in Table 33, and not the catalyst concentration itself. Support for this idea was provided by a simple control experiment: two different catalyst loadings from Table 33 were tested again, this time with the amount of 1- nitropropane adjusted so that the nitropropane/catalyst ratio would remain the same. These reactions gave the same enantioselectivity for the major (syn) product. These two catalyst loadings had afforded a significant difference in the enantioselectivities under the conditions in Table 33 (82% and 90% ee), and so the results here demonstrate clearly that 115 the nitropropane/catalyst ratio is the determining factor for the highly unusual selectivity trend associated with 82. Table 35. Cat. loading study with stable nitroalkane/catalyst ratio \ N02 N02 X mol% cat. 82 ——> N02 + /\/N02 Benzene,rt (x eq.) 80a X n-prNOZ/catalyst ratio %ee (syn) 20 100 92 5 100 92 The idea that the nitroalkane/catalyst ratio is determining the selectivities was further probed with a survey of the effects that the amount of l-nitropropane has on the reaction outcome. It would be expected that increasing the amounts of nitroalkane for a given catalyst loading should lead to increased inductions. Along these lines, a set of reactions was undertaken where catalyst loading and concentration were kept constant and the number of equivalents of nitroalkane was varied, from 2 to 30-fold excess with respect to the nitroolefin. This study served in two purposes: it was not only a control experiment for probing the importance of the nitroalkane/catalyst ratio, but also a survey for the determination of the optimum amount of l-nitropropane that should be used in the reaction. The results are presented in Table 36. 116 Table 36. Effect of nitroalkane equivalents (a) 20 mol% 01:3 QC 8 Q H H NJLN cr=3 g N\ w 00 p (0.1M) /N\ N02 + > /\/ N02 Benzene,rt 80a equivalents NMR yield% d.r. entry time(h) %ee syn %ee anti n-PrNO,_ (b) (syn/anti) 1 2 60 36 46/54 82 70 2 5 60 43 46/54 83 74 3 10 15 85 60/40 82 70 4 15 15 100 63/37 83 71 5 20 15 85 65/35 85 72 6 30 15 100 66/33 87 76 (a) Entries 1-3 were run for 60h. Entries 4—6 were run for 15h. All reactions gave complete conversions except entries 1,2. (b) Determined from the 'H-NMR of the crude reaction mixture using triphenylmethane as internal standard. The results from the study on the variation of the equivalents of nitroalkane were interesting both in terms of the effect on the diastereoselectivity and the enantioselectivity of the syn adduct. Concerning the diastereomeric ratio, there is a slight preference for the anti isomer at small excesses of the reagent (enties 1 and 2) that inverts to a 2/1 selectivity for the syn product when the reagent is used in more than 15 fold excess (entries 4 to 6). Concerning the enantioselectivity of the syn adduct, the induction stays at 82—83% ee for 2-15 equivalents of nitropropane, but for larger excesses of 20 and 30 equivalents an 117 increase at 87% ee was seen. The ee of the anti product also went up from 70 to 76%, with only the result in entry 2 showing a deviation from an otherwise smooth trend. The trends in the selectivities in general agree with the notion about the nitroalkane/catalyst ratio is directly proportional to the asymmetric induction. As a bonus from the optimization of the reaction through the above experiments, it was delighting to observe that the ee of the product can be increased by just adding more nitropropane to the reaction mixture. On the other hand, considering about the rate and the yield of the transformation, it became apparent that at least 10 equivalents of the nucleophile should be used in order to obtain good yields in a reasonable amount of time. The question of whether the catalyst concentration or the RNOzlcatalyst ratio is responsible for the selectivity trend has been answered, however the question of whether species B or C is operating is still pending, because the results of all these experiments to this point are in accordance with the involvment of either species. In other words, the RNOz/catalyst ratio should be expected to influence both species in the same direction, as illustrated in Figure 22. To distinguish more clearly between species B and C and to} determine whether the suspected self-association of the catalyst is intramolecular or intermolecular, non-linear studies were conducted. In non-linear studies the relationship between the enantiomerical purity of the product and that of the catalyst is surveyed. Usually a non-linear relationship indicates the existence of catalyst aggregates, while a linear one indicates that the catalyst functions as a monomer. For the purpose of this study, three different samples of scalemic catalyst were prepared by mixing (R) and (S) -82 to make catalyst samples that were 20%, 50% and 80% ee. The study was undertaken in reactions using both 2% and 20% catalyst, such that the possibility of aggregation at 118 different concentrations of the catalyst can be tested. In the beginning of the discussion it was mentioned that species B in Figure 22 should be favored at low RNOZ/catalyst ratios. The equivalents of nitropropane used in these experiments for 2% and 20% catalyst were 30 and 15 respectively, such that the conditions wouldn’t disfavor the formation of the suspected polymer in the second case. In other words, in case the high catalyst loading (20%) created a catalyst aggregate, the RN02/catalyst ratio used here was much smaller than in the reaction with 2% (75/1 versus 1500/1) so that it wouldn’t break up the aggregate and give a misleading linear plot. The results of the study are illustrated in Figure 23. “s..‘.‘.:... 2% e.g., 5...... stair... " ” Series 2: 20% cat" 15 eq. nitroalkane 100 90 80 70 60 < 50 -O- Seriesl + SerlesZ 30 20 « 10 . o . - c . - o 20 40 so so 100 120 %ee catalyst %ee syn product Figure 23. Plot of optical purity of 82 versus optical purity of syn-80a The non-linear studies in Figure 23 show that there is a linear relationship between the cc of the catalyst and the ee of the major diastereomer of the product 80a, both at 2% 119 and 20% catalyst loading. This indicates that even when the catalyst is used at high concentrations or loadings, it still functions as a monomeric species, even if the RNOz/catalyst ratio is much smaller. This study may not disprove the existence of aggregated catalyst, but it is a strong indication that the active species is a monomeric form of the catalyst 82. Consequently, the intramolecular H-bonded shown in Figure 22 as species C is the only scenario that fits all the data obtained so far. In case species C is operating than the data obtained from the concentration study in Table 34 make sense if the increase in the concentration leads to more efficient interaction between the catalyst and the substrates, and helps to break-up the itramolecular H-bond in the catalyst. Given the fact that excess l-nitropropropane can lead to increased asymmetric induction that presumably results from altering the catalyst structure through disruption of H-bonding leading to the conversion of species C to D (Figure 22), one could argue that other compounds that could mimic the H-bonging ability of l-nitropropane should in principle do the same. In other words, it would be expected that additives that could H- bond with the catalyst and break-up the self-association should promote the formation of species D just as nitropropane does, if our hypothesis is correct. This idea was intriguing not only because it would give us one more probe to test our hypothesis, but also because some additives might work even more efficiently than l-nitropropane affording even higher inductions and the identification of such additives would furthermore contribute to the optimization of the reaction. Along these lines, a series of additives were tested in the reaction of nitrostyrene and l-nitropropane using 20 mol% catalyst. The additives were chosen on the basis of mimicking the ability of l-nitropropane in forming H-bonds with the catalyst, and thus 120 small molecules that can function as H—bond acceptors were chosen. The results of this study are presented in Table 37. Table 37. Survey of additives in the reaction with catalyst 82 (a) 20 mol% CFa HJLH CF3 . H I N N \ \ No. CC / No2 (0.16M) N / \ _ No2 + N... 3233.... ' 3.. 10 eq. entry additive (eq.) yield% (b) d.r. ee% syn l - 100 66/34 88 2 PhNO2 (5) 83 58/42 92 3 THF (5) 62 63/37 89 4 MeCN (5) 83 64/36 92 5 DMSO (5) trace nd nd 6 EtOAc (5) 67 57/43 89 (a) Reactions run for 12 h. (b) Determined from the Il~I-NMR of the crude reaction mixture using triphenylmethane as internal standard. First of all, it was observed that all the additives slowed down the reaction, some to a small extent (PhNOz, MeCN) and others more significantly (THF, EtOAc) while only a trace of the product was obtained when DMSO was used. This is consistent with the presumtion that all the additives interact with the catalyst. In particular DMSO, as maybe 121 expected, seems to completely block the catalyst preventing its interaction with the substrates. Out of the five additives that were tested, both PhNO2 and MeCN afforded an increase in the ee from 88% to 92%. This small but real change suggests that these two additives interact with the thiourea of the catalyst shifting the equilibrium from the self associated catalyst C to the more selective complex D, in accordance with the proposed mechanism. It was also very interesting to observe that when acetonitrile is used as a solvent the yield and the selectivities drop significantly (51%, 62% ee), as is typically the case in thiourea catalysis when polar H-bonding acceptor solvents are employed. However when acetonitrile is used as an additive in small amounts it contributes to the enhancement of the ee. Finally, the hypothesis of catalyst self-inhibition through species C was further probed with temperature experiments. If, as proposed, catalyst self-inhibition is Operating as a factor in regulating the reaction selectivities, it would be expected that this inhibition should be favored at low temperatures, where the energy provided to the system is lower. As the temperature rises one would expect that the increased energy of the system should give to the molecules the necessary energy to overcome the entropic barrier of breaking an intramolecular H-bond in the catalyst which can lead to intermolecular H-bonds between the catalyst and the substrates. This scenario would lead to a counter inturtive outcome, lowering the temperature would lead to decreased asymmetric inductions. As a test of this hypothesis, the cc of the syn product was studied as a function of temperature in a range between -40 and 70°C. Unfortunately, at high temperatures the catalyst underwent decomposition according to the crude lH-NMR spectrum. Experiments at low temperatures provided quite interesting results. Lowering the temperature from 5 to 122 -10° C led to a sharp decrease in the induction from 84% to 59% ee. This is in accordance with internally H-bonded catalyst, because lowering the temperature, and thereby the energy provided to the system, should otherwise lead to increased inductions. The trend that was observed actually was not a gradual enantioselectivity decrease. As shown in Figure 24, the ee starts rising again when going to -20 and -30°C and at -40°C the induction is back to the initial levels. It seems that the internally H-bonded catalyst is maximized at -5 to -10°C, and Figure 24. Temperature study of the asymmetric addition 20% 82 N02 + N N02 N02 (01 M) (10 eq.) Benzene O 80 90 1,; 80* 7 . 60‘ {D \ 504 4o- % 08 (SW!) 30‘ 20- 10- A V r o 20 4b so 80 "Q at lower temperatures the increased selectivities from the reaction either through species D or C takes over and almost recovers the induction. 123 In any case, the sharp drop in the selectivity when lowering the temperature is not usually encountered in asymmetric catalysis when monomeric catalytic species is involved in the reaction, and the result that was observed here is an indication for some sort of intramolecular interaction, in agreement with the intramolecularly H—bonded species C. 5.3.2.4 Optimization of conditions and substrate scope From the studies described in the previous section, the effect of most of the reaction parameters has already been defined for catalyst 82 in the reaction of nitrostyrene with l-nitropropane. The next objective is then to achieve the best combination of these parameters so that the optimum balance between yields and selectivities can be identified. In the case of the asymmetric addition with catalyst 82, it was fortunate to observe that the unusual selectivity patterns described in the previous section were such that an increase in both the concentration and the amount of l-nitropropane favor an increase not only in the yields, but also in the diastereoselectivities and the enantioselectivities as well. Thinking about catalyst loading, tha data in Table 33 suggests that in order to make the reaction more selective small catalyst loadings have to be used. According to the the data however, loadings smaller than 20% afford incomplete conversions and low yields. Catalyst loading is apparently the only factor that has opposing effects on the yields and selectivities. On the other hand, since an increase in equivalents of nitropropane and higher concentrations favor the selectivities, it would be expected that using a large excess of nitropropane and high concentrations would help to overcome this issue of sluggish reactivity at 2 mol%, 1 mol% or even 0.5 mol% catalyst loadings. Therefore, further optimization was focused on the objective of “pushing” these reactions to give higher conversions and yields. The results are presented in Table 38. 124 Table 38. Optimization of reaction variables (a) CF3 00 3.25.. [£:/\/No2 /N\ No2 + ; No2 /\/ N02 benzene [C] prNO2 yield% ee% (c) entry‘ cat.load.% time (h) d.r. (c) nitrostyrene equiv. (b) (syn) 1 2 0.2M 20 16 23 84/16 94 2 2 0.2M 30 40 90 84/ 16 94 3 2 0.3M 30 40 100 80/20 93 4 l 0 .2M 30 40 66 87/ 13 95 5 0 .5 0 .2M 30 40 40 88/ 12 96 6 0.5 0.4M (neat) 30 40 40 80/20 94 (a) Reactions were run for 12h. (b) Determined from the 1H-NMR of the crude reaction mixture using triphenylmethane as intemal standard. (c) Determined by chiral HPLC. As can be seen from the data in Table 38, using 2 mol% catalyst affords only 23% yield of 80a and a lot of unreacted nitrostyrene was observed under the conditions in entry 1. However, this was solved by using more forcing conditions. Apparently increasing the amount of nitropropane, the reaction time and the concentration leads to an excellent NMR yield with the selectivities in the same high levels (entries 2 and 3). The same strategy was attempted for the reactions with 1 mol% and 0.5 mol% catalyst loadings, but 125 in these cases it was not possible to push the yields to a satisfactory level, even when neat conditions were employed (entries 4 to 6). From the above study summarized in table 38, the conditions in entry 2 were chosen as the optimum and then the scope of the asymmetric addition of l-nitropropane to nitroolefins was studied using these parameters. Various differently substituted nitrostyrenes were tested and the results are shown in Table 39. 126 Table 39. Substrate scope of the asymmetric addition of nitroalkanes to nitrostyrenes using catalyst 82 (a) NO MN N02 + r=l\/No2 2% mol. 82‘ l 2 7 N02 (0.2M) 30 eq benzene, rt R 800-1 overall yield% %ee entry Ar R adduct d.r. (0) yield% (b) syn (b) (syn) 1 Ph Et 80a 80 60 84/16 95 2 4-MeO-C6H4 Et 80b 60 47 85/ 15 94 3 2-MeO-C6H4 Et 80c 55 45 80/20 94 4 2-MeO-C6H4 Et 80c 94 72 80/20 94 5 4-Me-C6H4 Et 80d 72 58 80/20 94 6 2-Cl-C6H4 Et 80c 84 62 74/26 92 7 4-Cl-C6H4 Et 80f 75 62 83/17 92 8 4-Br-C..,H4 Et 80g 75 65 87/ 13 93 9 3-Br-C6H4 Et 80h 69 62 90/ 10 94 10 2-Br-C6H4 Et 80i 78 58 75/25 92 1 1 Ph n-Pr 80j 78 70 90/ 10 91 (a) Reactions run for 40h except entries 7 and 8 which were 17h and entry 4 which was 60h. All reactions were 0.2M in nitroolefin except enrty 4 which was 0.33M. (b) Isolated yields after column chromatographic purifiction. (c) Determined by the weight of the separated diastereomers except entries 1-4 which were determined by 'H-NMR. As can be seen from the data in the Table 39, the reaction is highly efficient for a range of substituted nitrostyrenes. Good isolated yields in the range of 69-94% of the adducts 80a-i can be obtained with only 2 mol% catalyst, demonstrating a catalyst with very efficient turnover ability. In the case of the electron rich methoxy-substituted styrenes 127 the rates of the reactions were slow when the optimum conditions were employed, and isolated yields of 47% and 45% were obatained for the syn adducts for the 4-MeO and the 2-MeO substrates respectively. This problem could be suppressed in the case of the 2- MeO adduct 80c (entry 4) by performing the reaction at a higher concentration (0.3M) for a prolonged time (60h). In this way an excellent overall yield could be obtained (94% overall) without any loss in the selectivities. The same didn’t work for the 4-MeO substrate, though, since its conversion remained the same (~70%) despite the efforts to push the reaction. Electron deficient chloro and bromo substituted nitrostyrenes reacted faster and gave mostly complete conversions of the starting material, some of them in short reaction times ( 17h for 4»Cl and 4-Br). On the other hand, electron neutral, withdrawing or donating groups do not seem to affect the selectivities. The diastereoselectivities are in the range of 74/26 tp 90/ 10 in favor of i the syn adduct, while the asymmetric inductions are all between 92-95%, demonstrating a highly efficient enantiodiscrimination with excellent consistency. On the other hand, the reaction with l-nitrobutane (entry 11) is consistently efficient, giving a slightly higher diastereomeric ratio along with 91% ee for the syn isomer, indicating that the reaction is not restricted to l-nitropropane, but could be potentially applied to higher nitroalkanes. Switching from trans-B-nitrostyrene to the more demanding trans-or-methyl-B- nitrostyrene proved disappointing. A reaction using 2 mol% catalyst was run for 2 days and afforded no product but only unreacted starting material (Scheme 32). 128 Scheme 32. Reaction with trans-a-methyl-B-nitrostyrene N rah/Y N02 + /\/N02 2%: mol. 82 O2 : N benzene, rt ph 02 2 days 0%, no reaction An effort to extend the scope to branched nitroalkanes turned out to be more promising. 2-Nitropropane features a secondary nitro group and thus is much less reactive than l-nitropropane. Addition of 2-nitropropane to nitrostyrenes have only been reported by Wang’s group52 and according to their protocol many (6-12) days are required for the completion of these transformations. When the reaction of 2-nitropropane with nitrostyrene was attempted using '10 mol% catalyst 82, a 55% yield of the adduct 83 was obtained with 68% ee (Table 40). Lowering the catalyst loading to 5% afforded a better induction (76%) which is in accordance with what has been seen for l-nitropropane, however, the increased induction is at the expense of the yield, although in this case the reaction was somewhat dilute (0.13M) and that might be a reason for the sluggishness. 129 Table 40. Reaction with 2-nitropropane using catalyst 82 Ph/V N02 + N02 Catalyst 82 32 \f benzene, rtr Ph N02 24h entry %catalyst equiv. 2-nitropropane NMR yield% %ee l (a) 10 27 55 68 2 (b) 5 55 27 76 (a) Reaction was run with 0.2M with respect to nitrostyrene. (b) Reaction was run with 0.13M with respect to nitrostyrene. Finally, an effort was made to extend the scope of the reaction to aliphatic nitroolefins. Nitroalkene 87 was synthesized according to a common procedure44b from butyraldehyde 85 through an addition of nitromethane to make 86 followed by an elimination of water using CuCl and DCC (Scheme 33). Reaction of 87 with l- nitropropane using 2 mol% catalyst 82 afforded a 66% conversion to the dinitro compound 88 in a clean reaction, according to the crude lH-NMR spectrum, as a 58:42 mixture of diastereomers with 42% and 28% ee’s for the major and the minor diastereomer respectively. The pruduct was not isolated though because most of it was lost during purification because of its volatility. 130 Scheme 33. Synthesis and reaction of the aliphatic nitroolefine 88 ? MeNOZ/NaOH 0“ DCC,CuCl EtOH/AcOH 7 «BK/N02 320, n WNoz 87% 42% 87 catalyst 82 N02 WNOz + /\/N02 (2 mol%) ; 3L3“) as 87 (0.35M) 30 eq. Dengue. rt 2 66% (NMR), d.r.=58/42 ee=42% (major) ee=28% (minor) Although the result is not up to the level of the reactions with nitrostyrene, the good conversion and the enantioselectivity of the major diastereomer suggest that a further optimization and effort to extend the reaction to other aliphatic nitroalkenes would be worthy. 5.3.2.5 Proposed Stereochemichal Model for the asymmetric addition using thiourea 82 According to the generalized mechanism for the role of bifunctional catalysts in the conjugate addition of nitroalkanes to nitroolefins that was illustrated earlier in the chapter (Scheme 29) the key reactive intermediate is believed to involve the DMAP moiety of the catalyst interacting with the deprotonated l-nitropropane as a cation/anion pair, along with the nitrostyrene which is held through H-bonding interaction from the thiourea group. A postulated mechanism that accounts for the observed stereochemical outcome with thiourea 82 is depicted in Figure 25. 131 Figure 25. Postulated stereochemical model NO observed /1/ 2 enantiomer ,.- N02 (28,3S)-808 (2R, 3Fl)-80a © ébfin ”’“o- Pb in ”k” N _ . H 0'0 I’ N+ W5 /N\ F30 3.0 ”*3 N / NQQ /\ / ‘N + \ %r Q A': favored B' _N~ (less strain) as S liq F30 Q @530. Q S 33 “J N o _ F3C H H H H N O'NO I ' N + H H , . N F3C H O. .0 / \ O O Q N + N N / N a / \ / —N\ ‘T \ \\ ’ DI Cl _N‘ \ N02 N02 (9%“ The. <28. 33...... (ZR, 3S)-803 © The absolute stereochemistry of the major enantiomer was determined by a comparison with the HPLC data and the optical rotation values of the corresponding compounds from the literature (for the known compounds 80a—c, 80c). The proposed stereochemical model is similar to the one for the BINAM derived bis-thiourea catalyst 52 that was developed for the asymmetric aza-Henry reaction and described in chapter 4. The major enantiomer (2R, 3R)—80a is believed to be derived via the intermediate A’, where 132 both components of the reaction orient themselves in the chiral pocket in such a way as to minimize the steric interaction with the (R)—BINAM moiety of the catalyst. The opposite enantiomer is obtained through B’, which is very unfavored due to the fact that both of the substrates experience unfavorable contacts with the binaphthyl core. The expected energy difference should be high, and it seems like a reasonable explanation for the excellent ee and the very efficient enantiodiscrimination. for the syn adduct. The same factors make intermediates C’ and D’, which afford the anti isomers, energertically unfavored and explain the diastereoselectivity in favor of the syn isomer, as well. In other words, the anti adduct is disfavored because both intermediates C’ and D’ involve steric interactions that are absent in A’. This model is in accordance with the higher diastereoselectivity that was observed for l-nitropropane and l-nitrobutane in comparison with nitroethane, since it would be expected that bulkier nitroalkanes would impose a higher energy cost for intermediate D’. Considering the cc of the minor diastereomer, which is determined by the energy difference between C’ and D’, it is smaller than the ee of the syn product and thus makes sense since C’ and D’ each has one negative steric interaction, and thus the dicrimination between these two diastereomeric transition states shouldn’t be as high as that for A’ versus B’. Again, as in the case of the aza-Henry bis-thiourea catalyst 52 (chapter 4), the above model is a working hypothesis that explains the stereochemical outcome. Given the conformational freedom that both the thiourea moiety and the DMAP group can experience other slightly different models could be proposed. The above model just represents a straight-forward representation of the reaction according to the assumptions described earlier. 133 5.4 SUMMARY AND CONCLUSION In conclusion, the thiourea catalyzed conjugate addition of nitroalkanes to nitroolefines has been studied both in its non-asymmetric and enantioselective version. Concerning the non-asymmetric Michael reaction, it was found that using 2 mol% of Schreiner’s thiourea 47 along with 2 mol% triethylamine can lead to the syn-selective synthesis of 1,3-dinitroalkanes in good yields. The successful employment of thiourea 47 to the addition of nitroalkanes to nitroolefins prompted the investigation of an asymmetric variant. As a result, the first highly efficient protocol for an organocatalytic direct addition of nitroalkanes to nitroolefines has been developed. The reaction is promoted by a novel bifunctional DMAP/thiourea catalyst and affords enantiopure syn-1,3—dinitro compounds with only 2% cat. loading. The asymmetric inductions in the products are remarkably consistent (92- 95% ee for the addition of l-nitropropane to nitrostyrene) and no drop-off is observed even when ortho-substituted styrenes are used. A unique selectivity trend was discovered for this reaction according to which, the less catalyst used, the higher the enantioselectivity of the reaction, this effect was correlated with the nitroalkane/catalyst ratio. It was demonstrated that an increase in this ratio leads to increasd inductions, presumably because the interaction of the catalyst with nitroalkane changes its structure in solution to a more selective species. This idea can be formulated as a self-deactivated catalyst, and a series of experiments were done to probe this. As a result of this unusual selectivity trend, ee’s up to 96% could be obtained with only 0.5% catalyst loading, demonstrating a catalyst that is remarkably selective. 134 HAPTER TOWARDS VAULTED BIARYL-DERIEVED THIOUREAS AND THE SYNTHESIS OF VANAM It was demonstrated in chapters 4 and 5 that 1,1’-binaphthyl-2,2’-diamine is an efficient chiral schaffold for thiourea mediated catalysis. The success of BINAM in inducing high enantioselectivities in the asymmetric addition of nitroalkanes to imines using catalyst 52 and to nitroolefines using catalyst 82 showed the potential of biaryl- based thiourea catalysts in a field dominated by cyclohexane diamine and Cinchona alkaloid-based organocatalysts. These results along with the work that has been done in our laboratory with the ligands VANOL and VAPOL (see chapter 1) prompted the consideration of catalysts 89- 90 that are based on the vaulted binaphthyl structural unit. Figure 26. Vaulted biaryl thioureas and diamines I <1 IZ 23: 12 ZI 0 TI on O CF3 Catalysts 89 and 90 feature a vaulted biaryl unit that would be expected to communicate stereochemical information closer to the reaction center, that is the thiourea 135 and/or DMAP moieties, and so their synthesis and application in these reactions — or other related transformations - would be highly desirable. The chiral diamine 91, which would be the starting material for the above catalysts, is not known in the literature. Therefore, the key issue in the synthesis of these novel catalysts is the synthesis of the diamine- analog of VANOL. Some preliminary efforts towards the synthesis of diamine 91 have been undertaken, and this chapter details the results of thess studies. 6.1 Examination of Buchwald amination as a key step in the synthesis of 91 Concerning the synthesis of arylamines, one of the most powerful protocols reported in the literature involve the reactions of aryl triflates (or bromides and chlorides) with primary or secondary amines under Pd catalysis, described by Buchwald and coworkers“. For example, the reaction of triflate 92 with benzylamine using a catalyst formed in situ from Pd(OAc)2 and ligand 94 affords the secondary arylamine 93 in 81% yield (Scheme 34)”. Scheme 34. Buchwald amination of triflate 95 1% Pd(OAc)2 0T1 NHBn 0 2°/ 16 O + BnNH2 o > O PU‘Bulz t-Bu NaOtBu 1.3” Toluene, rt 94 92 81% 93 It was reasoned that this kind of amination reaction could be used as a key step in the synthesis of 91 according to the sequence illustrated in Scheme 35. 136 Scheme 35. Proposed synthesis of 91 using Buchwald amination as a key step 0 Buchwald O I m'nat' - Ph OTf a ' '0” Ph NHBn deprotectlon ........ - Ph NH2 --------- ’ 95 96 91 A number of different amines could be coupled with VANOL triflate 95 — which should be easily obtained from VANOL — with benzyl amine being a good example of an amine that could be subsequently deprotected easily to give 91. The investigation of this stratgy started with the synthesis of triflate 95. The reaction of racemic VANOL with triflic anhydride in the presence of pyridine afforded a very good yield of the biaryl triflate 95, both on a small scale as well as on a 1g scale reaction (Scheme 36). Scheme 36. Synthesis of VANOL triflate o 0 ph OH Ph OTf Ph OH DCM Ph 0T1 116 I 95 I 150 mg scale: 91% 19 scale: 80% 137 Table 41. Initial efforts for the synthesis of 96 NHan O 10%. Pd(OAc)2 O on "98nd Ph NHBn OTf NaOtBu Ph NHBn toluene O 96 I ligand conditions result 94 rt, triflate added all at once 0% 96, triflate hydrolyzed BINAP 80°C, triflate added over 1h 0% 96, triflate hydrolyzed The first attempts to couple 95 with benzylamine using the conditions illustrated in Table 41 were disappointing. The reaction using ligand 94 at rt and adding the triflate all at once led to the hydrolysis of a big portion of 95 and the formation of free VANOL (Table 41). The hydrolysis of triflates during Pd catalyzed amination reactions is a common problem according to the reports of Buchwald and Hartwig". This hydrolysis is attributed to the rapid cleavage of the triflate by attack of NaOtBu (or other strong bases) on sulfur leading to the corresponding phenoxide. It is usually associated either with the electronic nature of the triflate (electron deficient triflates are prone to hydrolysis) or the strength of the base that is used. In the case of compound 95 there is no issue with the electronic nature of the substrate, but still it seems that the substrate is not stable under the given conditions. Hartwig has reported that slow addition of the triflate to the reaction mixture 138 can prevent the undesired hydrolysis“. Along these lines, an experiment was performed where 95 was added over a 1h period, but with no success (Table 41). A second way to avoid triflate hydrolysis is using a relatively weak base for the amine deprotonation step. Buchwald has reported an improved method using weaker bases such as CSZCO3 for the amination of halides and triflates that affords higher yields and improved functional group compatibility”. The use of C52CO3 was tested in the reaction of benzylamine with 95. The hydrolysis of 95 was completely prevented this time and no free VANOL was obtained, but the reaction didn’t afford any product either. All the starting material remained unreacted under the conditions shown in Scheme 37. Scheme 37 . An attempt for Buchwald amination of 95 using CszCO3 0 5% Pd(OAc)2 U 5.5% BINAP Ph on + _ on BnNH2 7 Ph NHBn 052003 Ph NHBn Toluene O 0 100002411 95 0%, NO REACTION 96 O At this point, it was suspected that the bulkiness of the VANOL triflate might be responsible for the failure of these attempts. Since 95 remains unreacted under forcing conditions even when hydrolysis is prevented, this probably means that the triflate is probably too bulky for the oxidative addition step, particularly when the ligand that is used for Pd also is considerably bulky. 139 To test this hypothesis it was reasoned that the coupling should be tried with the triflate 100 obtained from the mononer precursor of VANOL, 99. The result from the coupling of 100 should give information about weather the bulkiness of VANOL triflate is the reason for the failure of its coupling. To test whether the bulkiness of VANOL is responsible for its unreactivity in the Pd catalyzed coupling, triflate 100 was made in 80% yield from 99 and it was subjected in coupling reactions with benzylamine using different bases. The results are shown in Scheme 38. Scheme 38. Synthesis and amination of triflate 100 O OH "20 W G OTf O D O O 99 1 00 5%catalyst OTf base NHBn O + BnN H2 t O O Toluene O 100 101 entrty catalyst base T(°C) %yield 101 l Pd(OAc)2/BIN AP NaOtBu 80 0% 2 Pd(OAc)2/94 K3PO4 80 0% 3 Pd(OAc)2/BINAP CSZCO3 100 87% 4 Pd(OAc)2/94 CSZCO3 100 28% 140 The use of NaOtBu afforded the same result as for the attempted coupling of 95. A big amount of 100 was hydrolyzed and 99 was obtained as a major product from the reaction in entry 1. The use of K3PO4 afforded no reaction and recovery of unreacted 100. The use of CszCO3 though afforded the desired product 101 in excellent yield when BINAP was used as a ligand (entry 3) and with a lower yield when 94 was employed (Scheme 34). These two last results in entries 3 and 4 confirm that the triflate 95 is presumably too bulky for the Pd catalyzed coupling under the above conditions. The fact that the monomer 100 can afford an excellent yield whereas the VANOL triflate gives no reaction implies that the formation of the intermediate of the oxidative addition that features Pd coordinated both by BINAP and VANOL triflate is probably disfavored due to steric congestion. 6.2. Examination of a synthesis of 91 through oxidative coupling of arylamines A second pathway that was examined for the synthesis of 91 involves the use of oxidative coupling between substituted anilines. The synthesis of 91 through oxidative diaryl coupling in fact has been attempted before in the literature by Kocovsky’s group, but without success”. When Kocovsky employed the free aniline 102 in an oxidative coupling using CuCl2 he obtained carbazole 103 as the only product in 25% yield. The driving force for the formation of compound 103 is the loss of ammonia. Kocovsky proposed that the loss of ammonia is not occurring through the formation of 91 as an intermediate, but instead, via a competing pathway that involves N-N coupling of 102 and a subsequent 3.3-rearrangement. 141 Scheme 39. Oxidative coupling of 102 using CuCl2 NFI [Illa EEBG ph l NH2 CUClg. 2H20 / BnNH2 Ph 0 MeOH IAnisole P“ rt 102 103 Specifically, through the oxidative coupling of a number of substituted aniline derivatives and control experiments, he postulated the mechanism shown in Scheme 40 to account for the formation of the resulting diamines and carbazoles. Scheme 40. Postulated mechanism for the CuCl2 mediated oxidative coupling of substituted anilines so“ coco 104 \ 105 Cu(ll) NH2 --/---+ Ii H sllli] lVHg /lllii 107 142 According to this mechanism, naphthylamines like 104 can undergo a Cu(II) mediated oxidative coupling either affording directly 5] (C-C coupling) or 105 through coupling at nitrogen. Intermediate 105 can undergo a 3,3-sigmatropic rearrangement to give 106, which subsequently can tautomerize to 51 or alternatively lead to the carbazole precursor 107, that can release ammonia to afford 108. For 2-naphthylamine 104 the major product was the diamine 51 while only a trace of carbazole was detected (~l%). For the coupling of other aniline derivatives though the percentage of carbazole formation was much higher and in the case of the coupling of 102, as mentioned before, only carbazole was obtained. This mechanism was supported by the fact that the Bucherer-type reaction of 2-naphthol with hydrazine that is expected to proceed through intermediate 105, was found to readily afford 5159. Additionally, when the isolated diamines like 51 were subjected to the condition of the oxidative coupling no carbazole was obtained, which implies that the carbazole byproducts are not produced by loss of ammonia from the diamine, but instead, via the alternative N-N coupling pathway. The same control experiment was not performed with VANAM since no diamine was produced or isolated from the coupling of 102, but the authors presume that its behavior should be similar to that of 10458. Concerning the synthesis of VANAM, it was reasoned that if compound 102 was substituted at nitrogen then the formation of the carbazole byproduct should be disfavored since the ammonia loss would not be an issue anymore. This way, even in the case of N-N coupling the 3,3-rearrangement would hopefully result in the preferential formation of the desired diamine. This synthetic route using the oxidative coupling as a key step would add one more key step to the sequence, which is the resolution of the racemic diamine 96. 143 Along these lines, substrate 101 was tested in the oxidative coupling using CuCl2 and following the conditions reported by Kocovsky for the corresponding reaction with 102. The results, shown in Scheme 41, were rather disappointing. Koockovsky’s conditions (r.t.) afforded no reaction and complete recovery of the starting material. An attempt to push the reaction by heating at 100°C also did not produce a reaction. ph NHBn CuClz, 2H20 I BnNH2 O Scheme 41. Attempted oxidative coupling of 101 using CuCl2 Ph NHBn MeOH IAnisole 101 96 ! temperature (°C) result 25 no reaction 100 no reaction It was suspected that the failure of the above reaction might be a result of the bulkiness of the substrate 101. The position of coupling is ortho, ortho- disubstituted by a phenyl group and a benzyl group, and that might impose a considerable energy barrier for the coupling of the two units. In this case, it was envisioned that the solution may be the employment of substrates that bear smaller N-substituents. 144 Scheme 42. Synthesis and attempted oxidative coupling of 111 Pd(OAc) H Ph on BINAP 2 Ph NA CI + WNHQ CI C52CO3 0 Toluene O 100 109 100°C 111 entry scale time % catalyst equiv. 109 % yield 1 300 mg 12h 5 1.2 14 2 300 mg 2d 10 2.4 60 3 1 g 2d 10 2.4 57 Ph A O CUClz_ 2H20 / BHNHZ Ph D .. N O MeOH/Anisole,rt V Ph 0 ”A 0%, No REACTION O 111 112 The allyl-protected substrate 111 was the compound of choice for testing this hypothesis. Compound 111 was synthesized from a Pd catalyzed coupling of 100 with allylamine 109. This amination was a little more problematic than the one with benzylamine, since allylamine is a volatile compound and easily evaporates from the reaction mixture when using high temperatures, thus affording low yields (entry 1). This was solved when a larger excess of 102 was employed and higher catalyst loading was used to increase the rate. In this way a good yield of the desired product was obtained even when the reaction was scaled up to one gram of material. 145 It was disappointing to see that 111 did not give any product when its oxidative coupling was attempted (Scheme 42). Similar to 101, the attempted reaction with 111 afforded a clean reaction mixture where surprisingly absolutely nothing had happened. The failure of the CuCl2 oxidation method turned our attention to alternative, and more forcing conditions for the oxidative coupling reaction. In specific, oxidative coupling under neat conditions using high temperature and air was considered. This has been a standard method for the synthesis of the chiral ligand VAPOL in our laboratories’, and we thought it might turn out to be successful for the arylamine coupling as well. The coupling under neat conditions was tried both for 101 and 111. The solids were put in a flask and subjected to high temperature for 3.5 hours under the flow of air. The results are shown in Scheme 43. Perhaps not surprisingly, the sensitive terminal double bond in 111 seems to be responsible for the considerable amount of baseline material observed in the oxidation of 111. Concerning 101, the only product observed after 3.5 h of heating was Scheme 43. Attempted oxidation under neat conditions H P“ N\/\ neat, air 0 H O \ Ph NV + basel'ne 1 18000 P" ”A material 3.5h O 1“ 0 112.0% Ph NHBn . O 0 "eat” 8‘" Ph NHBn P“ vah 180°C Ph NHBn + 3.5 h 0 O 101 I 96. 0% 113, 32% 146 the imine 113 resulting from the oxidation of the benzyl moiety. These were the last attempts concerning this particular approach. 6.3. Attempts to synthesize 91 using the Bucherer reaction The last approach that was considered for the synthesis of 91 was the use of the Bucherer reaction, that is, the direct displacement in one step of the hydroxy groups with ammonia under forcing conditions”. The Bucherer reaction has been reported to transform 2—naphthol 110 and the VANOL monomer 99 to the arylamines 104 and 102 respectively”, via the mechanism shown in Scheme 44. Scheme 44. The Bucherer reaction for the synthesis of 104 and 102 CH (NH4)2SO3.H20‘ NH2 ””3 / H20 110 1500,1011 104 Ph 0H 0 P“ 0 NH3 / H20 200°C, 2d 99 42% 102 CH NaHSO3 O NH3 NH = —— 0. H20 11° soaNa 303Na Z I M l2 9) I U) 0 co 2 I m 104 SOaNa 147 Despite the usefulness of the name reaction in the synthesis of simple naphthyl amines, there is no precedent for the synthesis of 1,l’-biaryl-2,2’-diamines using the above method. In fact, BINOL (120) has been reported to afford the monoaminated product 121 when subjected to the reaction conditions, but no BINAM was produced“. The difference in the reactivity of BINOL compared to its monomer 110 has been explained on the basis of an intramolecular H-bonding that stabilizes the monoaminated product 121 and “shuts down” any further reactivity, stopping the reaction and preventing the second hydroxyl group from getting displaced by ammonia (Scheme 45). Scheme 45. Bucherer reaction of BINOL I I OH (NH412503.H2C3 I I OH 0” NH3 / H20 ”“2 200°, 5d 91 °/o 120 121 Despite the unfavorable precedent with BINOL it was decided that an effort to test the VANOL (and VAPOL) ligand in the Bucherer conditions would be worthwile. VANOL and VAPOL thus were put in an aqueous solution of ammonia and heated at high temperatures in the presence of NaHSO3. The results are shown in Table 42. 148 Table 42. Bucherer reactions of VANOL and VAPOL O \ ’3') I”) ’0 O aq. NH3 Ph OH NaH303 Ph OH + O , 11" sq i, \ ’3’) 3’). ,0. 116, VANOL 91 1, VAPOL 117 %yield %yield entry ligand scale T(°C) time (h) 91 or 117 118 or 119 1 116 217mg 200 60 0 so 2 116 2g 200 40 0 25 3 1 270mg 220 24 0 0 As can be seen, the reaction of VANOL 116 on a small scale (entry 1) produced compound 118 in 50% yield with no trace of the desired 91. When the reaction was performed on a 2g scale though, the isolated yield dropped to 25% despite the almost complete conversion of the starting material. Finally, for the VAPOL ligand (entry 3) no reaction occurred at 220°C. Not surprisingly, the reactivity of VANOL resembles that of BINOL and the suspected intramolecular H-bond controls the reaction outcome here as well. 149 6.4. Future Work In a recent report by Cho and coworkers it was shown that diaryl hydrazides 125- 127 prepared from the Pd(0) catalyzed coupling reactions between aryl bromides 122-124 and di—tert-butyl hydrazidoformate (BocNHNHBoc) can undergo acid-catalyzed rearrangement to form 2,2’-diamino-l ,1’—binaphthyls, according to the sequence shown in Scheme 46”. Scheme 46. Cho’s synthesis of BINAM derivatives BocNHNHBoc cat. HCl CO Pd2(dba)3’xa"p"°: NBoc EtOH/80°C R NH2 122 (R=H) 125, 81% 51, 75°/ 123 R=Br 125, 37% 128, 313/o 124 (R=Me) 127, 65% 129, 81% This type of rearrangement has been known for a long time to proceed either thermally or with acid catalysis for 105, giving BINAM as a major product. For example, heating of 105 in boiling EtOH affords an 82% yield of 51 and a small amount (~15%) of the carbazole 108 (Scheme 47)”. It seems that EtOH promotes the rearrangement by acting as a general acid catalyst. 150 Scheme 47 . Thermal rearrangement of 105 for the synthesis of BINAM ”NH EtOH/80°C CO NH2 # “.3 105 + NH 108 (15%) 00 51 (82%) It was envisioned that this synthetic sequence could be potentially used for the synthesis of VAN AM 91. An N-protected diaryl hydrazide 131 could hopefully rearrange to afford the desired product 91 without significant production of carbazole. If Kocovsky’s proposed mechanism for the diamine versus carbazole formation is correct and the carbazole is produced through ammonia loss from the 3,3-rearrangement pathway intermediate 106, then one would expect that in case of 131 the ammonia loss should be prevented if the rearrangement occurred before the deprotection of nitrogen. The mechanism for the reaction in Scheme 47 is not clear in every detail, but it could be argued that Scheme 48. Proposed synthesis of 91 through hydrazide rearrangement 2 Ph 130 Br BocNHNHBoc Pd2(dba)3/xanphos 08003 Ph I NBoc EtOH/80°C Ph Ph 0 NBoc 1h Ph i O 9, O E NH2 NH2 131 151 avoiding the use of HCl in the rearrangement of 131 and performing the reaction under milder conditions (thermally, using only EtOH) might allow for the reaction to occur with the Boc groups still on, thus eliminating the issue of ammonia loss. Compound 130 could be obtained from the triflate 100 through the intermediacy of a boronate ester 132 according to the method depicted in Scheme 49. This chemistry has been developed by Huffman’s group and is known to afford very good yields of aryl halides“. For example, 2-naphthyl and l-naphthyl bromides can be obtained in 77% and 84% overall yields from the starting triflate respectively. The reaction for the particular triflate 100 is not known, but it is expected that it will hopefully show similar behavior. Scheme 49 Proposed synthesis of 130 Ph OTf Ph 3' Ph Br 0 . 0 ‘° O 100 132 130 i. pinacolborane, PdClz, dppl, dioxane, NEt3; ii. CuBrz, MeOH, H20 In addition, the direct coupling of triflate 100 with BocNHNHBoc could be examined as an alternative way of accesing 131 in a fewer number of steps. Cho’s method only concerns naphthyl bromides, but an extension of the methodology to include aryl triflates might be feasible. In this case the synthesis of 91 would be shorter and thus more attractive . 152 EXPERIMENTAL PR ED RES Experimental procedures for Chapter 2 6-Methylnapthoic acid 22 0 COOH AIC|3 Me 00 21 ~ 5 % 22 2-Furoic acid (120 g, 1071 mmol, 1.0 equiv) was added portionwise to a 3-necked flask containing toluene 21 (1000 mL), which was immersed in an ice bath and fitted with a mechanical stirrer. A bubbler was attached before AlCl3 (300g, 2249 mmol, 2.1 equiv) was added in spatula portions over a 2 h period, maintaining a steady evolution of gas. The flask was removed from the ice and slowly heated to 60 °C and then stirred at this temperature for 18 h. The solution was then poured into an ice-cold HCl solution (550 ml of cone HCl in 2 litres of H20). This was then heated overnight (18 h) at 60 °C, to dissolve any solids. The two resulting layers were separated, and the organic layer was washed with H20 (2 x 300 ml) before being heated with a 1.0 M NaHCO3 solution (1000 mL) at 55 0C for 1 h. This was repeated once more and the combined aqueous layers (2000 mL) were heated to 55 °C before Ba(OH)2.8 H20 (315.0 g, 1000 mmol) was added and the resulting solution stirred for 2 h. The yellow precipitous solution was filtered and the red/brown aqueous layer collected and split into 3 x 1 litre fractions. Using pH paper these solutions were titrated with cone HCl to a pH of 7.0. The resultant yellow precipitate was 153 filtered and the green aqueous filtrate was further titrated with conc. HCl to a pH of 5.0- 6.0. A paler yellow precipitate was filtered and collected, TLC analysis indicated that it was the desired product. The purity of the solid varied, with a solid orange contaminant being present. The material was crystallized to purity from either hot benzene or EtOAc to deliver bright yellow crystals of 6-methylnapthoic acid, (16.7 g, 85.6 mmol, 8%). 1H NMR (300 MHz, MeOD) 6 2.39 (s, 3H), 7.30-7.38 (m, 2H), 7.58 (s, 1H), 7.85 (d, 1H, J = 8.5 Hz), 8.00 (dd, 1H, J: 7.0, 1.2 Hz), 8.68 (d, 1H,J= 9.0 Hz). Diisopropyl-6-methynapthamide 24 O OH 0 CI 0 N(iPl’)2 socrz (iPr)2NH, Et3N _ Me 00 Benzene, DMF 00 CH2012. 0°C 00 Me 74 % from 20 Me 22 23 24 To a flask containing 6-methylnapthoic acid 22 (8.27 g, 44.46 mmol, 1.0 equiv) under N2 was added benzene (18.5 mL, 2.4 M) and the resulting slurry stirred for 5 min. SOCl2 (3 .5 mL, 48.11 mmol, 1.1 equiv) was added followed by the dr0pwise addition of DMF (1.5 mL, 19.36 mmol, 0.44 equiv). The resulting clear red solution was allowed to stir for 4 h, after which time the solution was concentrated on the rotary evaporator, removing any excess SOClz. The benzene/DMF solution was taken on into the next step. The above reaction was assumed to have progressed in quantitative yield. 154 CH2C12 (10 mL) was added to a flask containing the acid chloride under N2. This solution was stirred for 5 min before being added via canula to a flask containing a stirring solution of (i-Pr)2NH (7.48 mL, 53.38 mmol, 1.2 equiv) and NEt3 (7.44 mL, 53.38 mmol, 1.2 equiv) in CHZCI2 (70 mL, 0.5M in 23) under N2. After stirring overnight, H20 (50 mL) was added to the reaction, the layers separated and the aqueous layer extracted with CHzCl2 (2* 100 mL), and the combined organic fractions washed with 2M HCl (aq) (2*100 mL), sat NaHCO3 (aq) (2*100 mL) and dried with NaZSO4. Concentration under reduced pressure and column chromatography (10% EtOAc2hexane) gave diisopropyl-6- methynapthamide (32.9 mmol, 8.85 g, 74 % for the two steps) as a white solid; 1H NMR (300 MHz, CDC],) 6 1.01 (d, 3H, J = 6.5 Hz), 1.05 (d, 3H, J = 6.5 Hz), 1.63 (d, 3H, J = 7.0 Hz), 1.69 (d, 3H, J = 7.0 Hz), 2.48 (s, 3H), 3.62-3.54 (m, 2H), 7.21-7.42 (m, 3 H), 7.60 (s, 1 H), 7.69-7.73 (m, 2 H); 13C NMR (300 MHz, CDCls) 5 20.6, 20.6, 20.8, 21.6, 45.9, 51.0, 121.2, 124.7, 125.3, 127.1, 127.4, 127.8, 128.9, 133.7, 136.0, 136.4, 170.0, 1 sp3 C not located; IR (neat) 3053m, 2976m, 16225, 1332m, 12655, 710$, 704$ cm"; mass spectrum m/z (% rel. intensity) 269 [M‘] (16), 169 (100), 141.0 (43), 115 (l 1), 88 (17), 84 (16), calcd for m/z 269.1780, meas 269.1777. Anal calcd for C18H230N: C, 80.26; H, 8.61; N, 5.20. Found: C, 80.02; H, 8.73; N, 5.18; White solid (EtOAc) mp = 140-142°C; R,=0.19 (10% EtOAczhexane). 155 Diisopropyl 6-methyl-2-(2-phenylallyl)naphthalen-l-amide 25. o N(iPr)2 1) s-BuLi O N(iPr)2 2) Mg/C2H4Br2 O 00 3) Bromo Methyl Styrene ~75 °/ 0‘ Me 0 Me 24 25 s-BuLi (6.43 mL, 8.99 mmol, 1.1 equiv) was added dropwise over 1 min to a solution of diisopropyl-6-methylnapthamide (2.2 g, 8.17 mmol, 1.0 equiv) in THF (80 mL, 0.08 M in 24) under N2 at —78 °C. The resulting orange solution was stirred for l h before the canular addition of the ethereal MgBr2 solution, which was prepared in situ as described below. Dibromoethane (2.1 mL, 24.51 mmol, 3.0 equiv) was added over a 1 h period to a flask, fitted with a condenser, containing magnesium tumings under N2. The resulting reaction was stirred for 2 h before being added via canular to a solution containing the ortho-lithiated napthamide. The colour of the solution changed from orange to pale green and it was then stirred for a further 15 min before being allowed to warm to rt over a l h period. The solution was then cooled to —78 °C, followed by the dr0pwise addition of or- bromomethyl styrene and then allowed to warm to room temperature overnight. The reaction was quenched with sat NH4C1 (100 mL), the layers separated and the aqueous fraction extracted with EtOAc (2*100 mL), washed with brine (2*50 mL) and dried with NaZSO4. Concentration under reduced pressure and column chromatography (10% EtOAczhexane) gave the desired diisopropyl 6-methy1-2-(2-phenylallyl )naphthalene-l - amide (2.35 g, 6.12 mmol, 75%) as a pale yellow foam. 'H NMR (300 MHz, CDCl,) 6 156 1.03 (d, 3H, J = 6.5 Hz), 1.11 (d, 3H, J = 6.5 Hz), 1.69 (d, 3H, J = 6.5 Hz), 1.81 (d, 3H, J = 6.5 Hz), 2.51 (s, 3H), 3.54—3.72 (m, 2H), 3.88 (d, 1H,J= 16.5 Hz), 4.12 (d, 1H, J: 16.5 Hz), 5.07 (d, 1H, J = 1.5 Hz), 5.64 (s, 1H), 7.26-7.75 (m, 10H); 13C NMR (300 MHz, CDC13) 5 20.5, 20.6, 21.0, 21.3, 21.5, 38.0, 46.1, 50.9, 115.8, 124.7, 126.0, 126.9, 127.0, 127.3, 127.5, 128.0, 128.3, 128.6, 131.1, 132.3, 134.7, 135.3. 140.6. 145.9, 169.4; IR (CDC13) 3055, 2986, 2936, 1622, 1265 cm" ; mass spectrum m/z (% rel. intensity) 385 [M’] (10), 285 (25), 269 (30), 249 (30), 226 (15), 169 (70), 141 (30), 88 (100), 71 (75), 57 (65), 51 (100), calcd for m/z 385.2406, meas 385.2416.R,=0.45 (20% EtOAc:hexane). 7-Methyl-phenylphenanthren-4-ol 26 o N(iPr)2 ”0 Ph 0 MeLi. THF 0Q 00 Me so % Me 25 25 MeLi (5.52 mL, 5.52 mmol, 2.2 equiv) was added dropwise over a 5 min period, to a solution of diisopropyl 6—methyl-2-(2-phenylallyl)naphthalen-l-amide (0.966 g, 2.509 mmol, 1.0 equiv) in THF (50 mL, 0.05 M) at -—78 °C, followed by stirring for l h before allowing the reaction to warm to room temperature overnight. The reaction was quenched with sat NH4Cl (60 ml), the layers separated and the aqueous layer extracted with EtOAc (3 x 60 mL). The combined organic fractions were washed with brine (2 x 30 mL) and dried with NaQSO4. Concentration under reduced pressure and column chromatography (10% EtOAc:hexane), resulted in isolation of the desired 3-phenyl-8-methylphenanthrol 157 as a brown solid (0.602 g, 2.00 mmol, 80%). 1H NMR (300 MHz, CDCl,) 6 7.23 (d, 1H, J = 1.5 Hz), 7.28 (s, 1H), 7.42 (d, 1H, J = 7.5 Hz), 7.51 (t, 3H, J = 7.5 Hz), 7.73 (t, 6H, J = 10 Hz), 9.53 (d, 1H, J: 8.5 Hz); l3C NMR (300 MHz, CDCl,) 6 21.3, 112.0, 118.6, 119.7, 127.1, 127.2, 127.4, 127.8, 128.1, 128.2, 128.3, 128.8, 132.7, 134.9, 135.6, 138.7, 140.2, 154.4, 1 sp2 C not located; IR (CDC13) 3566, 3055, 2988, 1205, 710, 706 cm"; mass spectrum m/z (% rel intensity) 284 M+ (100), 241 (10), 184 (50), 141 (60), 115 (20), 57 (30), calcd for m/z 284.1201, meas 284.1197. Brown solid (hexane), mp = 146-147°C, R,=0.36 (20% EtOAc:hexane). 7 ,7 ’-Dimethyl-2,2’-diphenyl-(3,3’-biphenanthrene)-4,4’-diol 11 Ho Ph 0 0 180-190°C _ H0 Ph DO air HO 0 Ph 26 00 11 A round bottom flask containing 3-phenyl-8-methylphenanthrol (0.49 g, 1.33 mmol) was fitted with condenser. The flasks were then heated in an oil bath to 185 °C for 42 h. At 20 h interval, the black solid that was forming was dissolved with EtOAc, to ensure good mixing, and the polymer was filtered through celite, before the mixture was heated again. The reaction was monitored by TLC and lH-NMR. Work-up of the reaction involved dilution with EtOAc, filtration of insoluble polymeric material and column chromatography (4% EtOAc:pentane) leading to the isolation of name ( 164 mg, 0.438 158 mmol, 33.5%) as a slightly impure orange solid. Some starting material and a third compound was recovered - not characterized. 'H NMR (300 MHz, CDC13) 6 2.58 (s, 6H), 6.65-7.64 (overlapping m, 22H), 9.63 (d, 2H, J = 8.5 Hz); ”C NMR (300 MHz, CDC13) 6 21.3, 115.7, 118.1, 123.1, 126.7, 127.0, 127.5,12803, 128.08, 128.68, 128.70, 128.8, 128.9, 133.0, 134.9, 136.1, 139.8, 141.1, 153.2; IR (neat) 3484w, 3026w, 2916w; mass spectrum m/z (% rel intensity) 566 M” (100), 384 (13), 283 (83), 244 (28), 154 (18), 105 (12), calcd for m/z 566.2246, meas 566.2245. NOTE: The reaction using air flow is slower and generates more byproducts (S)-7 ,7 ’-Dimethyl-2,2’-diphenyl-(3,3’-biphenanthrene)-4,4’-diol 11 O CuCl, (-)-Sparteine O HO Ph ultrasound HO Ph “0 P“ MeOH/DCM HO Ph 0 o ( ‘ 11 S)-11 (Typical procedure for deracemization of 7,7’-dimethyl VAPOL) To a 50 mL round bottom flask was added 12 mL of MeOH, 64.7 g of CuCl (0.65 mmol) and 0.30 g of (-)-sparteine (1.30 mmol). The flask was sonicated in an ultrasonic bath with air sparged into the solution. The temperature was maintained at or below 25 0C by the addition of ice to the water in the bath. The flask was sealed with a septum after 45 minutes and then purged with argon for 60 minutes in the ultrasonic bath. The green 159 Cu(II)-sparteine solution was transferred via cannula to the solution of racemic 7,7’- dimethyl VAPOL (0.22 g, 0.38 mmol) in 50 ml of CH2C12 in a 250 m1 round bottom flask which was had previously been purged with argon and sonciated for 60 minutes. The combined solution was sonicated for 10 minutes and then the flask was removed from ultrasonic bath, covered with aluminum foil and stirred at room temperature for 6 hours under a positive pressure of argon maintained by slow flow through the head space of the flask. The reaction was quenched by adding 5 mL of concentrated HCl and then stirring for 10 minutes. The two layers were separated. The lower CHZCl2 layer was collected, the upper MeOH/1120 layer was washed with 3*30 mL of CHZClz. The combined CHzCl2 layer was dried over MgSO4. Upon removal of solvent, the crude reaction mixture was loaded to silica gel column and eluted with a 1:1 mixture of CHzClzl hexane to give 0.17 g (0.30 mmol, 79% yield) 26 as a white solid. The optical purity of 26 was found to be of 2 99 % by HPLC on a Pirkle D—Phenylglycine column with a 75:25 mixture of hexane/isopropanol (260 nm, flow rate: 2 mL/min). Under these conditions the retention time of (R)-25 was 14.8 min and that of (S)-25 was 19.8 min. 160 A typical experimental procedure for the reaction of aldimine with silyl ketene acetals. To VAPOL (0.063 g, 0.11 mmol) and Zr(O-i-Pr),.iPrOH (0.05 mmol) in toluene (0.9 mL) was added l-methylimidazole (0.06 mmol) in toluene (0.1 mL) at room temperature. The mixture was stirred for l h at the same temperature. An aldimine (0.25 mmol in 1 mL) toluene solution was added to the catalyst and the mixture was stirred for an additional 5 min. Then the silyl ketene acetal (0.3 mmol) was added. The mixture was stirred for 15 h at room temperature. Aqueous N aHCO3 was added to quench the reaction. The mixture was extracted with CH2C12 (2 x 10 mL). The combined organic layer was concentrated to give the crude product. The crude product was treated with THF and aqueous 1N HCl (10:1) at 0°C for 30 min. Then the reaction was quenched with aqueous NaHCO3 and extracted with ethyl acetate (2 x 10 mL). The combined organic layer was washed with brine and concentrated. The desired product was obtained by chromatography on silica gel. The optical purity was determined by chiral HPLC analysis. Methyl 3-(2-hydroxy-phenylamino)-3-phenyl-2,2-dimethyl-propionate 4. OH PhAN + >_<0Me —’ NH 0 OH Ph OMe 2 3 4 The spectral data for this compound are the same as those previously reported for this compound (ref. 8). 'H NMR (300 MHz, CDCl,): 6 = 1.20 (s, 3H), 1.23 (s, 3H), 3.67 (s, 3H), 4.53 (brs, 1H), 4.85 (brs, 1H), 5.29 (s, 3H), 6.37 (dd, J = 7.8, 1.1 Hz, 1H), 6.52 (dt, J = 7.0, 1.2 Hz, 1H), 6.59 (dt, J = 7.8, 1.1 Hz, 1H), 6.70 (d, J = 7.6 Hz, 1H), 7.20 - 161 7.27 (m, 5H). The optical purity was determined by chiral HPLC analysis (Daicel Chiralpak AD, Hexanes/i-PrOH = 9/1, flow rate 1.0 mL/min): R, = 9.56 min (major enantiomer), R, = 12.71 min (minor enantiomer). tert-Butyl 3-(2-hydroxy-3,5-dimethylphenylamino)-3-phenyl-propionthioate 27a. 31"“ NH O * StBu 27a White solid; Rf = 0.32 (9/1 hexanes/ethyl acetate); mp 90 — 92°C; lH NMR (300 MHz, CDC13): 6 = 1.41 (s, 9H), 2.05 (s, 3H), 2.17 (s, 3H), 2.85 (dd, J = 4.9, 14.8 Hz, 1H), 2.96 (dd, J = 8.8, 15.1 Hz, 1H), 4.17 (s, 1H), 4.68 (dd, J = 4.9, 8.8 Hz, 1H), 5.36 (s, 1H), 6.17 (s, 1H), 6.39 (s, 1H), 7.27—7.30 (m, 5H); 13C NMR (75 MHz, CDCl,): 6 = 15.8, 20.9, 29.6, 48.5, 51.7, 56.7, 115.0, 122.2, 122.7, 126.3,127.2,128.5,129.3,134.2,l41.8,141.9, 198.6; MS (EI) m/z (relative intensity): 327 (9) [M’], 227 (18), 226 (100), 148 (8), 136(5), 131(2), 91 (12), 77 (5); Anal. Calcd. for C20H25NO3: C, 73.37; H, 7.69; N, 4.28. Found: C, 71.19; H, 7.63; N, 4.09 (failed); HPLC (Daicel Chiralpak AD, hexanes/i-Pr= 9/1, flow rate = 1.0 mL/min): Rt = 11.7 min (minor enantiomer), R1 = 15.5 min (major enantiomer), (5)-VAPOL used as ligand. 162 tert-Butyl-3-(2-hydroxy-3,5-dimethylphenylamino)-3-(1-naphthyl)-propionthioate fl... NH O O * S‘Bu Q White solid; R, = 0.31 (8/2 hexanes/ethyl acetate); mp 122 — 124°C; 'H NMR (300 27b. MHz, CDCl,): 6 = 1.45 (s, 9H), 2.00 (s, 3H), 2.21 (s, 3H), 2.95 (dd, J = 9.3, 14.8 Hz, 1H), 3.12 (dd, J = 4.1 , 14.8 Hz, 1H), 4.74 (br, 1H), 5.09 (br, 1H), 5.64 (dd,J = 4.1, 9.3 Hz,1H), 6.13 (s, 1H), 6.35 (s, 1H), 7.38 — 7.65 (m, 4H), 7.76 (d, J = 8.2 Hz, 1H), 7.90 (dd,J = 1.4, 8.2 Hz, 1H), 8.23 (d, J = 8.4 Hz, 1H); 13C NMR (300 MHz, CDC13): 6 = 15.8, 20.9, 29.7, 48.6, 51.2, 52.3, 113.3, 121.1, 122.3, 122.4, 123.1, 125.4, 125.5, 126.3, 127.8, 129.0. 129.8, 130.4, 133.9, 134.8, 137.1, 140.7, 198.4; HPLC (Daicel Chiralpak AD, hexanes/i- PrOH = 9/1, flow rate = 1.0 mL/min): R, = 12.9 min (minor enantiomer), R, = 16.5 min (major enantiomer), (S)-VAPOL used as ligand. 1.63 tert-Butyl 3-(2-hydroxy-3,5-dimethylphenylamino)-3-(4-chlorophenyl)-propionthioate fl... NH O * StBu 14c. (31 27c Yellow oil; Rf = 0.09 (9/1 hexanes/ethyl acetate); 1H NMR (300 MHz, CDC13): 6 = 1.48 (s, 9H), 2.13 (s, 3H), 2.24 (s, 3H), 2.89 (dd,J = 5.4, 15.0 Hz, 1H), 2.99 (dd, J = 8.4, 15.0 Hz, 1H), 4.75 (m, 1H), 6.19 (s, 1H), 6.45 (s, 1H), 7.30 — 7.34 (m, 6H); l3C NMR (300 MHz, CDC13) 15.7, 20.8, 29.6, 29.7, 48.7, 51.6, 55.9, 114.6, 122.2, 122.7, 127.8, 128.7, 129.6, 132.9, 134.1, 140.6, 141.6,198.3 ppm; IR (neat): 3410, 2964, 2922, 1672, 1601, 1516, 1491, 1456, 1385, 1199, 1151, 1091, 1030, 1014, 985, 827, 783 cm"; LRMS (FAB’) m/z (relative intensity): 393 (33, 37Cl) [NF], 39] (76, 35Cl) [M’], 260 (20), 197 (20), 135 (60), 109 (24), 95 (40), 83 (40), 69 (60), 57 (100); HRMS (FAB+) Calcd for C2,H26NOZSC14391.1373, found 391.1375; HPLC (Daicel Chiralpak AD, hexanes/i-PrOH = 9/1, flow rate = 1.0 mL/min): R, = 17.3 min (minor enantiomer), R, = 26.2 min (major enantiomer), (S -VAPOL as ligand). 164 Experimental Procedures for Chapter 4 General procedure for the preparation of the catalysts 52, 55. 3,5-Bis-trifluoromethylphenyl isothiocyanate (8.12 mmol, 1.48 mL) was added to a solution of the appropriate diamine (4.06 mmol) in 7 mL THF at rt and the mixture was stirred overnight. The volatiles were evaporated and the crude mixtures were purified by crystallization. Catalyst 52 043 1152311310 OW”? .N 001:: THF, 0°C to rt 96% (FD-51 (Fl)- 52 CF3 (R)-BINAM was used (1.26 g, 4.00 mmol). Purified by crystallization from dichloromethane/hexanes to afford 3.22g (3.89 mmol, 96%) isolated yield. White crystals, m.p. = 120-122°C. 1H NMR (CDC13, 300 MHz): 6 = 7.09 (d, J = 8.7 Hz, 2H), 7.26 (t, J: 6.9 Hz, 2H), 7.48 (m, 4H), 7.54 (m, 2H), 7.67 (br s, 4H), 7.81 (d, J = 8.7 Hz, 2H), 7.95 (d, J = 8.3 Hz, 2H), 8.10 (d, J = 8.7 Hz, 2H) ppm; 13C NMR (CDCl3, 300 MHz): 6 = 119.2, 122.8 (q,J= 272.9 Hz), 124.6, 125.3, 125.6, 126.7, 127.5, 127.6,128.5,129.7,131.7 (q,J = 33.7 Hz), 132.3, 132.8, 134.4, 139.1, 180.4 ppm; IR (neat) v = 3265, 1688, 1498, 980 cm"; HRMS (ESI) calcd for C38H23N4FUSZ: 827.1179, found 827.1172; [alzsD = +103.8 (c = 1.0, acetone). 165 Catalyst 55 2 eq. C48 F30 N” CF3 © 1 Q _,.NH2 01:3 ‘ ..u u CF3 E l E l H NH2 THF, 0°C to n “TN cps (5'56“ 74% s (S.S)-55 Synthesized according to the general procedure using 468 mg (4.10 mmol) (S,S)- 1,2-cyclohexyldiamine, purified by crystallization from EtOAc/hexanes to afford 1.99 g white crystals (3.03 mmol, 74%). IH NMR (300 MHz, CDC13) 6 = 1.35 (br s, 4H), 1.81 (br s, 2H), 2.20 (br s, 2H), 4.38 (br s, 2H), 7.07 (br s, 2H), 7.69 (s, 2H), 7.81 (s, 4H), 8.12 (br s, 2H); 13C NMR (300 MHz, CDCl,) 6 = 24.4, 31.7, 59.4, 119.6, 122.7, 124.0, 132.9, 138.59, 180.54. The data agreed with the reported literature 3“" Catalyst 53 S 00 0C 3 CO H .- (R) BINAM Cl/u\0| / CHCla N:(C:=: (9)-BIN AM “A“ - 3 .‘N= : \N N choa I H20 00 550/0 . H \n/H 90% ICC S 0 (m-57 (Rum-53 Synthesized using 365 mg (1 mmol) (R)-2,2’-bis(isothiocyonato)-2,2’-binaphthyl diamine” and 281 mg (1 mmole) (R)-l,l’-binaphthyl-2,2’-diamine to afford 351 mg (55%) of white crystals (acetone/hexanes). lH-NMR (CDCl3, 300 MHz): 6 = 6.68 (d, J = 8.7 Hz, 2H), 6.82 (s, 4H), 6.87 (s, 2H), 7.02 (d, J = 8.4 Hz, 2H), 7.15 (t, J = 6.9 Hz, 2H), 7.40 (q, J = 6.9 Hz, 6.9Hz, 4H), 7.57 (s, 2H), 7.67 (t, J = 6.9 Hz, 2H), 7.76 (d, J = 8.1Hz, 166 2H), 7.90 (d,J = 8.1Hz, 2H), 8.02 (d, J = 9.3 Hz, 2H), 9.41 (d, J = 9.3 Hz, 2H) ppm; l3C- NMR (CDC13, 300 MHz): 6 = 119.8, 120.8, 123.1, 125.4, 125.5, 125.9, 127.1, 127.2, 127.8, 128.2, 128.6, 128.8, 129.3, 130.2, 131.1, 132.2, 132.3, 132.5, 133.7, 135.7, 177.6 ppm. IR (KBr) 3154, 1595 cm"; MS (FAB’): m/z (%): 653 [M+1-I]+ (18), 619 (4), 460 (5), 307 (40), 252 (45) 154 (100), 136 (84), 102 (38); mp. 230-231°C. Catalyst 54 \ WN N=C =3 histamine fill OxNzc :8 O flN H)-57 ( H)-54 Synthesized according to the general procedure using 0.5 g (1.3 mmol) (R)-2,2’- bis(isothiocyonato)-2,2’-binaphthyl and 302 mg (2.71 mmol) histamine to afford 540 mg (0.91 mmol, 100% from 57) of an off-white solid. lH-NMR (DMSO-d6, 300 MHz): 6 = 2.55 (br. s, 4H), 3.47 (br. s, 4H), 6.62 (br. s, 2H), 7.07-7.17 (m, 4H), 7.40 (m, 4H), 7.73 (m, 4H), 7.95 (t, J = 9.3, 4H), 8.61 (br. s, 2H), 11.71 (br. s, 2H); l3C-NMR (DMSO-d6, 300 MHz): 6 = 27.0, 44.5, 55.4,126.1,126.2,126.7,127.2,128.3, 128.7, 131.9, 133.1,135.1, 136.4, 181.3; two carbons are not located. MS (FAB’): m/z: 591 [M+H]” (17), 307 (14); IR (neat) 3150, 1595, 1490 cm"; HRMS (FAB’) calcd for C32H3INSSZ: 591.2113, found 591.2117. 167 *— ‘M.~‘.p- -, -. General procedure for the preparation of catalysts 56-61a Following the general procedure for 52 and 55, the appropriate isothiocyanate (2 equiv.) was added to a solution of (R)-BINAM in THF (0.6M in BINAM) and the mixture was stirred overnight at rt. The volatiles were evaporated and the crude mixtures were purified by column chromatography on silica gel (always using 30% acetone in hexanes). Catalyst 56a, compound 56b. (Po-56a (50%) (H)-56b (50%) Catalyst 56a Synthesized from (R)—BINAM (300 mg, 1.05 mmol) and isolated as a white solid (277 mg, 0.50 mmol, 50%), mp. = 128-130°C; 'H-NMR (CDC13, 300 MHz): 6 = 6.12 (d, J = 7.2 Hz), 6.92 (t, J = 7.2 Hz, 4H), 7.03 (t, J = 7.5 Hz, 2H), 7.35-7.28 (m, 4H), 7.58-7.53 (m, 4H), 7.89 (d, J = 8.7 Hz, 2H), 8.0l(d, J = 2.7 Hz, 4H), 8.04 (d, J = 2.7 Hz, 4H), 8.30 (br. s, 2H); three exchangeable H’s do not show up. l3C-NMR (CDC13, 300 MHz): 6 = 125.30, 125.33, 126.4, 127.23, 127.29, 128.0, 128.1, 129.2, 129.5, 132.31, 132.33, 135.0, 135.6, 179.7; one C not located. IR (neat) 3159, 1593, 1535 cm"; MS (FAB*): m/z (%): 555 [M+H]*, (26), 462 (17), 307 (30), 154 (100), 136 (60); (1250 = +600 (c = 1.0, dichloromethane). 168 Compound 56b. White solid, 50% isolated (210 mg, 0.5 mmol) , m.p. = 108- 110°C; lH-NMR (CDCl3, 300 MHz): 6 = 6.46 (d, J = 9MHz, 2H), 6.83 (d, J = 7.8 Hz, 1H), 7.07-6.92 (m, 4H), 7.31-7.24 (m, 3H), 7.13 (td, J = 1.5, 6.9 Hz, 1H), 7.51-7.41 (m, 1H), 7.52 (br s, 1H), 7.85 (d, J = 8.7 Hz, 2H), 7.96 (d,J = 8.4 MHz, 1H), 8.36 (d, J = 9 Hz, 1 H), 8.40 (br s, 1H); three exchangeable H’s do not show up. '3C-NMR (CDC13, 300 MHz): 6 = 111.5, 118.3, 122.5, 122.9, 124.5, 125.6, 125.8, 126.0, 126.7, 126.7, 126.9, 127.2, 127.4,127.9,128.1,129.6,132.3,132.4,133.4,135.2,135.3, 142.7,179.0 ppm; 2 CS not located. MS (FAB+): m/z (%): 420 [M+H]+ (40), 307 (30), 154 (100), 136 (65); IR (neat) 3352, 3186, 1626, 1510; a251,: +24.2 (c = 1.0, dichloromethane). Catalyst 57a. ONOZ 3 N02 00 .0 CO NH2 SCN 11% ,,NH _,,N so 2 T so “do N02 (FD-57a (1 00%) Synthesized from (R)—BINAM (300 mg, 1.05 mmol) and isolated as a white solid (644 mg, 1.00 mmol, 100%). 'H-NMR (CDCl3, 300 MHz): 6 = 6.69 (d, J = 9 Hz, 4H), 7.18 (d, J = 8.7 Hz, 2H), 7.31 (t, J = 8.0 Hz, 2H), 7.53 (t, J = 8.0Hz, 2H), 7.82 (d, J = 9 Hz, 4H), 7.98 (d, J = 8.1 Hz, 2H), 8.06 (s, 4H) ppm; 4 exchangeable H do not show up. 13C-NMR (CDC13, 300 MHz): 6 = 112.8, 122.1, 124.7, 126.3, 126.6, 126.9, 127.1, 127.6, 128.5, 132.3, 133.1, 136.4, 142.9, 145.9, 180.0 ppm; MS (FAB*): m/z: 645 [M+H]", 507. 460, 307, 154, 136; IR (neat) v: 3410, 1702, 1594, 1500, 1388, 1268 cm“; m.p. = 134- 136°C; [a]25D = -56.7 (c = 1.0, acetone). 169 Catalyst 58a Cl (3)538 Cl (1 00%) Synthesized from (R)-BINAM (300 mg, 1.05 mmol) and isolated as a white solid (692 mg, 1.00 mmol, 100%), mp. = 116-120°C; lH-NMR (CDC13, 300 MHz): 6 = 6.50 (s, 4H), 7.04-7.03 (m, 2H), 7.30-7.22 (m, 4H), 7.55-7.48 (m, 2H), 7.74 (s, 2H), 7.88 (d, J = 9Hz, 2H), 8.00 (t, J = 9 Hz, 4H), 8.57 (br. s, 2H); ”C-NMR (CDC13, 300 MHz): 6 =179.8, 137.5, 135.2, 134.9, 132.12, 132.11, 128.9, 128.5, 128.2, 127.2, 127.0, 126.6, 126.3, 125.1, 123.6; IR (neat) 3434, 1644 cm"; MS (FAB‘): m/z (%): 693 [M+H]+ (l9, 3*35C1, 1*37Cl ), 691 [M+H]+ (8, 4*”Cl), 530 (1 l), 154 (100), 136 (60); [a]25D = +106.8 (c = 1.0, dichloromethane). Catalyst 59a, compound 59b. OMe Catalyst 59a. Synthesized from (R)—BINAM (300 mg, 1.05 mmol) and isolated as a white solid (104 mg, 0.17 mmol, 17%). lH-NMR (CDC13, 300 MHz): 6 = 3.77 (s, 6H), 6.18 (d, 6.9 Hz, 4H), 6.47 (d, 6.9 Hz, 8H), 7.27-7.56 (m, 8H), 8.02-8.15 (m, 4H) ppm; 13C- NMR (CDCl3, 300 MHz): 6 = 55.5,114.9,125.2,126.4,127.3,127.6,127.9,128.1,128.2, 170 132.3, 135.9, 158.9, 180.7 ppm; 3 CS not located. MS (FAB‘): m/z: 615 [M+H]*, 492, 307, 154; mp. = 124-125°C; IR (neat) v = 3163, 2953, 1508, 1244, 860 cm "; [alzso = +15.3 (c = 1.0, acetone). Compound 59b. Synthesized from (R)-BINAM (300 mg, 1.05 mmol) and isolated as a white solid (224 mg, 0.50 mmol, 50%); mp. = 114-116°C; lH-NMR (CDC13, 300 MHz): 6 = 3.68 (s, 3H), 6.45-6.38 (m, 4H), 6.81 (d, J = 8.4 Hz, 1H), 7.15-7.09 (m, 1H), 7.29-7.22 (m, 3H), 7.37 (br. s, 1H), 7.50-7.45 (m, 1H), 7.70 (d, J = 8.7 Hz, 1H), 7.84 (d, J = 8.7 Hz, 2H), 7.94 (d, J = 8.4 Hz, 1H), 8.02 (d, J: 9.0 Hz, 1H), 8.15 (br. s, 1H), 8.49 (d, J = 8.7 Hz, 1H) ppm; 2 exchangeable H’s do not show up; '3C-NMR (CDC13, 300 MHz): 6 = 55.6, 111.6, 115.5, 123.3, 125.6, 125.9, 126.3, 127.0, 127.1, 127.6, 127.8, 128.3, 128.43, 128.47, 128.5, 130.0, 132.6, 132.7, 133.7, 133.8, 135.8, 143.2, 158.7, 179.9, 122.7, 118.7 ppm; IR (neat) 3350, 3184, 3055, 2961, 2837, 1620, 1531 cm"; MS (FAB*): m/z (%): 450 [M+H]+ (98), 416 (20), 284, (100), 154 (70), 136 (50); [31250 = +662 (0 = 1.0, dichloromethane). Compounds 61a, 61b. NH2 1-Naphthyl-NCS ..~NH2 THF, rt (H)-61b (55%) Compound 61a. Synthesized from (R)-BINAM (300 mg, 1.05 mmol) and isolated as a white solid (32 mg, 0.05 mmol, 5%), not pure enough to characterize completely. IH- 171 NMR (CDC13, 300 MHz): 6 = 5.91 (d, J = 7.2 MHz, 2H), 6.85 (t, J = 8.1 MHz, 2H), 7.59- 7.14 (m, 16 H), 7.73 (dd, J = 8.7, 17.4 MHz, 4H), 7.88 (t, J = 8.4 MHz, 4 H); 2 exchangeable H’s do not show up. R, = 0.17 (30% acetone in hexanes); MS (FAB"): m/z (%): 655 [M+H]+ (40), 512 (37), 436 (20), 391 (18). Compound 61b. Synthesized from (R)-BINAM (300 mg, 1.05 mmol) and isolated as a white solid (259 mg, 0.55 mmol, 55%); mp. = 126-128°C; ‘H-NMR (CDC13, 300 MHz): 6 = 3.1 (br. s, 2H), 6.42 (d, J = 7.2 Hz, 1H), 6.71 (d, J = 8.4 Hz, 1H), 6.76 (d, J = 8.7 Hz, 1H), 6.92 (t, J = 7.8 Hz, 3H), 7.01 (t, J = 7.2 Hz, 1H), 7.21 (t, J = 8.1 Hz, 3H), 7.56-7.43 (m, 3H), 7.63 (d, J = 6.0 Hz, 1H), 7.66 (d, J = 9.0 Hz, 1H), 7.76 (d, J = 8.1Hz, 1H), 7.81(d,J= 8.1Hz, 1H), 7.93 (d, J: 8.1 Hz, 1H), 8.01 (d,J= Hz, 1H), 8.53 (d, J = 9 Hz, 1H), 8.63 (br. s, 1H) ppm; l3C-NMR (CDC13, 300 MHz): 6 = 110.9, 117.8, 122.0, 122.3, 122.7, 124.3, 125.00, 125.05, 125.5, 125.8, 126.5, 126.6, 126.9, 127.1, 127.6, 127.9, 128.03, 128.05, 128.4, 129.4, 129.5, 130.8, 132.1, 132.2, 133.0, 134.1, 135.3, 142.3, 179.9 ppm; 2 C’s not located. MS (FAB+): m/z (%): 470 (60) [M+H]", 307 (20), 284 (40); IR (neat) 3350. 2959, 1620, 1531 cm"; [a]25D = +4.7 (c = 1.0, dichloromethane). (R)-N-(l-(2-Aminonaphthalen-l-yl)naphthalen-2-y1)acetamide 63. CO CO ° NH A020, ACOH: “k 2 _,\NH2 DCM .\‘NH2 00 (FD-51 (FD-63 To a solution of (R)-(+)-1,1’-Binaphthyl-2,2'-diamine (284 mg, 1.0 mmol) and AcOH (0.6 mL, 10 mmol) in 10 mL of dried CH2C12 was added acetic anhydride (104 111., 172 1.0 mmol) at 0 °C under N. The resulting solution was stirred overnight at room temperature, then 2N NaOH aqueous solution was added until pH z 7. The reaction mixture was extracted by CH2C12 (3 x 50 mL) and the combined organic phases were washed with saturated brine and dried over MgSO4. The solvent was removed under reduced pressure and the crude product was purified by flash chromatography (ethyl acetate/hexane = 2/1) to afford a colorless oil in 77% yield (0.25 g, 0.77 mmol). lH-NMR (CDC13, 300 MHz): 6 = 1.85 (s, 3H), 6.91-7.42 (m, 8H), 7.81‘8.03 (m, 4H), 8.58 (d, J = 9.0 Hz, 2H). The data agreed with the reported literature‘wb. (R)-N-(l-(2-(Dimethylamino)naphthalen-l-yl)naphthalen-Z-yl)acetamide 64 O O CO “k formaldehyde CO “k ACOH THF (FD-63 85% (F064 N-(l-(2-Aminonaphthalen-1-yl)naphthalen-2-yl)acetamide 63 (0.25 g, 0.77 mmol) and aqueous formaldehyde (37%, 0.75mL, 9.0 mmol) were combined in 10 mL of THF and stirred for 15 min. NaBH3CN (200 mg, 5.3 mmol) was added, followed 15 min later by AcOH (1.0 mL). The resulting solution was stirred for 4 h at room temperature, then 1N NaOH aqueous solution was added until pH z 7. The reaction mixture was extracted by CHzCl2 (3 x 50 mL) and the combined organic phases were washed with saturated brine and dried over MgSO4. The solvent was removed under reduced pressure and the crude product was purified by flash chromatography (ethyl acetate/hexane = 1/5) to afford a brown powder in quantitative yield (272 mg, 0.77 mmol). 1H-NMR (CDC13, 300 MHz): 173 6 = 1.88 (s, 3H), 2.58 (s, 6H), 6.95 (d, J = 8.7 Hz, 1H), 7.12-7.55 (m, 6H), 7.84-7.80 (m, 4H), 8.49 (d, J = 9.0 Hz, 1H). 1 exchangeable H does not show up. The data agreed with the reported literature”. (R)-1-(2-(Dimethylamino)naphthalen-l-yl)naphthalen-2-amine 65 o “k 4MHC| EtOH NH2 ..N\ \ _ 0O \ 00"“ (FD-64 (H)-65 To a solution of N-(1-(2-(dimethylamino)naphthalen-1-yl)naphthalen-2- yl)acetamide 64 (0.18 g, 0.51 mmol) in 15 mL of EtOH was added 4M HCl (6 mL). The resulting solution was stirred for overnight at room temperature, then 1N NaOH aqueous solution was added until pH 2: 7. The reaction mixture was extracted by CHZCI2 (3 x 50 mL) and the combined organic phases were washed with saturated brine and dried over MgSO4. The solvent was removed under reduced pressure and the crude product was purified by flash chromatography (ethyl acetate/hexane = 1/10) to afford a colorless oil in 93% yield (148 mg, 0.47 mmol). 'H-NMR (CDC13, 300 MHz): 6 = 2.59 (s, 2Me), 7.0-7.29 (m, 7H), 7.47 (d, J = 9.0 Hz, 1H), 7.74-7.91 (m, 4H). 174 (R)-l-(3,5-Bis(trifluoromethyl)phenyl)-3-(l-(2-(dimethylamino)naphthalen-1- yl)aaphthalen-Z-yl)thiourea 66 CF13 CF3 so we go . (j NH2 CF3 “in CF3 .\N_ t CC (FD-65 (me To a solution of 1-(2-(dimethylamino)naphthalen-1-yl)naphthalen-2-amine 65 (36 mg, 0.12 mmol) in 2 mL of dried CHZCl2 was added 3,5-bis(trifluoromethyl)phenyl isothiocyanate (22 mg, 0.132 mmol) at 0 °C under N2. The resulting solution was stirred for overnight at room temperature. The reaction was concentrated in vacuo and the residue was purified by flash chromatography on silica gel (ethylacetate/hexane = 1/ 10) to afford a slight yellow solid in 91% yield (64 mg, 0.11 mmol); 1H NMR (300 MHz, CDC13): 6 = 2.59 (s, 6H), 6.90 (d, 1H, J = 7.5 Hz), 7.09 (t, 1H,] = 7.5 Hz), 7.26 (m, 2H), 7.36 (s, 2H), 7.41 (s, 1H), 7.56-7.50 (m, 4H), 7.71 (d, 1H, J = 8.5 Hz), 7.82 (d, 1H, J = 8.0 Hz), 7.98 (d, 2H, J = 9.0 Hz), 8.06 (d, 1H, J = 8.5 Hz), 8.37 (s, 1H); MS (FAB*): m/z (%): 355 (100), 252 (30), 157 (37), 140 (47), 123 (37), 73 (60). The data matched the reported literature‘o". 175 General procedure for the asymmetric aza-Henry reaction (Table 20): 20% catalyst (FD-52 N’B°° 40%EtaN HN’BOC | = NO Ar) 10 eq RCHZNOZ Ph/yxr 2 toluene, -35°C R 41 42, 433-1: A flame—dried round bottom flask was loaded with 0.2 equiv. of catalyst 52 (0.2 mmoles, 160 mg) and 1 equiv of imine 41a-k (1 mmol). The compounds were dissolved in 4 m1 of toluene. MeNOZ was added (10 equiv., 0.52 mL) at -35°C and after 5 minutes Et3N was added (0.4 equiv, 56 uL). The mixture was stirred at —35°C for 17-36 hours and then the volatiles were evaporated. The crude mixture was purified by column chromatography on silica gel (always 20% acetone in hexanes) to afford products 42, 433- R. For all the optimization studies of the aza-Henry reaction detailed in Chapter 4, the experimental procedures are the same as for the optimized reaction (described before), with slight modifications that are clarified in each Table. Whenever a % conversion is mentioned, that means the product was not purified, and the conversion was calculated from the integration of product peaks versus the imine C(=N)H proton, given that the only species observed in the crude lH-NMR was the catalyst, the desired product and imine, whenever some of it remained unreacted. tert-Butyl (R)-2-nitro-l-phenylethylcarbamate 42 NHBoc N02 (FD-42 176 According to the typical procedure, imine 41a (1 mmol, 205 mg) and MeNO2 were stirred for 36 h and converted to the product 42 (146 mg, 0.55 mmol, 55%) as a white solid. HPLC analysis (Chiralpak OJ-H, hexane/2-PrOH 95/5, flow rate = 1.0 mL/min, l = 210 nm): tr (minor) = 32.3, tr (major) = 36.8 min; [a]259 = -18.0 (86% ee, 0 = 1.0, CHC13); mp. 107-108 °C; 1H-NMR (300 MHz, CDCl3): 6 = 1.40 ppm (s, 9H); 4.64-4.81 (m, 2 H), 4.80 (br s, 1H), 5.34 (br s, 1 H), 7.37-7.23 (m, 5 H); l3C-NMR (300 MHz, CDC13): 6 = 28.3, 52.9, 78.9, 80.7, 126.4, 128.7, 129.2, 137.0,154.9 ppm. Palomo’s group (ref. 27) reported the same absolute configuration (R) for 42 with a negative Optical rotation, while Takemoto’s group (ref. 37) assigned the same absolute configuration (R) to a positive optical rotation. While it is not clear which assignment is right, we assigned the absolute stereochemistry of 42 following Palomo’s data (which agrees with the majority of the literature reports), and subsequently assigned the stereochemistry of the rest products (including 43k and 431) using 42 as a reference. 1-(p-Chlorophenyl)-2-nitroethyl carbamic acid t-butyl ester 43a NHBoc N02 CI (R)-43a Following the general procedure 239 mg imine were used (1 mmol) and compound 43a was obtained as a white solid in 62% yield (0.62 mmol, 182 mg). The ee of the product was determined by HPLC using a Daicel Chiralpak OJ-H column (n-hexane/i- PrOH = 97/3, flow rate = 1 mL/min, tr (minor)=68 .6 min; tr (major)=74.l min, [alzsD = - 44.0 (c = 1.0, CHClg); 85% ee; lH-NMR (300 MHz, CDCl3): 6 = 1.47 (s, 9H), 4.72—4.90 177 (m, 1H), 4.73 (dd, J = 5.0, 12.6 Hz, 1H), 5.40 (br s, 2H), 7.28—7.30 (m, 2H), 7.34—7.41 (m, 2H); l3C-NMR (300 MHz, CDC13): 6 = 28.2, 52.3, 78.7, 80.9, 127.8, 129.4, 134.6, 135.6, 154.8. The data agreed with the previously reported literature”. 1-(m-Chlorophenyl)-2.nitroethyl carbamic acid t-butyl ester 43b. NHBoc No2 Cl (H)-43b Following the general procedure, 239 mg imine were used (1 mmol) and compound 43b was obtained as a white solid in 53% yield (161 mg, 0.53 mmol). The cc of the product was determined by HPLC using a Daicel Chiralpak AD column (n- hexane/i-PrOH = 97/3, flow rate = 1 mL/min, tr (minor) = 23.2 min; tr (major) = 33.6 min, [31250 = -24.4 (c=1.0, acetone); 91% ee; lH-NMR (300 MHz, CDCl;,): 6 = 1.46 (s, 9H), 4.80-4.90 (br s, 1H), 4.72 (m, 1H), 5.42 (br s, 1H), 5.70 (br s, 1H), 7.21—7.35 (m, 4H); l3C—NMR (300 MHz, CDC13): 6 = 28.2, 52.2, 78.6, 80.9, 124.6, 126.7, 128.8, 130.4, 135.0, 139.2, 154.9. MS (FAB‘): m/z (%): 303 [M+H]* (3, 37C1), 301 [M+H]+ (10, ”Cl), 245 (90), 184 (90) 154 (82); IR (neat): v: 3370, 2980, 1686, 1556, 1368, 1165 cm"; m.p. 98—100°C. 1-(o-Chlorophenyl)-2-nitroethyl carbamic acid t-butyl ester 43c. NHBoc N02 Cl (9)-43C 178 Following the general procedure, 239 mg imine were used (1 mmol) and compound 7d was obtained as a white solid in 61% yield (179 mg, 0.61 mmol). The cc of the product was determined by HPLC using a Daicel Chiralpak AD column (n—hexane/i- PrOH = 99/1, flow rate = 1 mL/min, tr (major) = 69.1 min; tr (minor) = 92.3 min, [a]25D 2 +190 (0 = 1.0, acetone); 74% ee; lH-NMR (300 MHz, CDC13): 6 = 1.47 (s, 9H), 4.84 (m, 2H), 5.78 (m, 2H), 7.31-7.46 (m, 4H); ”C-NMR (300 MHz, CDCl3): 6 = 28.2, 50.6, 77.5, 80.8, 127.5, 128.0, 129.9, 130.3, 132.6, 134.4, 154.6. MS (FAB*): m/z (%): 303 [M+H]+ (3, 37C1), 301 [M+H]+ (10, 35C1), 245 (90), 184 (90) 154 (82); IR (neat): 3355, 2982, 2935, 1712, 1685, 1555, 1367, 1253 cm"; mp. 106-108°C. 1-(p-Bromophenyl)-2-nitroethyl carbamic acid t-butyl ester 43d. NHBoc N02 Br (FD-43d Following the general procedure, compound 43d was obtained as a white solid in 50% yield (172 mg). The cc of the product was determined by HPLC using a Daicel Chiralpak OJ-H column (n-hexane/i-PrOH = 95/5, flow rate = 1 mL/min, tr (minor) = 51.1 min; tr (major) = 54.9 min, 78% ee; 1H-NMR (300 MHz, CDC13): 6 = 1.47 (s, 9H), 4.75 (m, 1H), 4.76—4.90 (br s, 1H), 5.40 (br s, 2H), 7.23 (d, J = 6.9 Hz, 2H), 7.53 (d, J = 6.9 Hz, 2H); 13C-NMR (300 MHz, CDC13): 6 = 28.2, 52.2, 78.6, 80.9, 122.7, 128.1, 132.3, 136.1, 154.7; MS (FAB‘): m/z (%): 347 [M+H]“ (10, 8'Br), 345 [M+H]+ (10, 79Br), 289 (60), 228 (65), 154 (100); IR (neat): 3338, 2981 , 2935, 2363, 1687, 1527, 1166 cm"; m.p. = 140-142. 179 tert-Butyl (R)-2-nitro-l-(p-methoxyphenyl)ethylcarbamate 43c. NHBoc N02 (FD-43:9 According to the typical procedure, 235 mg imine were used (1 mmol) and the product 43c was obtained in 49% (145 mg, 0.49 mmol) as a white solid. HPLC analysis (Chiralpak OJ-H, hexane/EtOH 95/5, flow rate=l .0 mL/min, 1:210 nm): tr (minor) = 65.8, tr (major) = 72.0 min; [a]25D = -31.0 (89% ee, c = 1.00, CHCl3); 1H-NMR (300 MHz, CDC13): 6 = 1.47 ppm (s, 9H), 3.82 (s, 3 H), 4.69 (dd, J = 12.0, 5.3 Hz, 1 H), 5.36 (m, 1H), 4.83 (s, 1H), 5.42 (s, 1 H), 6.92 (d, J = 8.6 Hz, 2 H), 7.25 (d, J = 8.6 Hz, 2H) l3C-NMR (300 MHz, CDC13): 6 = 28.3, 52.4, 55.3, 78.9, 80.6, 114.5, 127.7, 129.0, 154.9, 159.8 ppm. The data agreed with the previously reported literature”. tert-Butyl (R)-2-nitro-1-(o-methoxyphenyl)ethylcarbamate 43f. NHBoc N02 0 I (H)-43f According to the typical procedure, 235 mg imine were used (1 mmol) and the product 43f was obtained in 40% yield (116 mg, 0.4 mmol) as a white solid. HPLC analysis (Chiralpak OJ-H, hexane/2-PrOH= 90/10, flow rate=1.0 mL/min,): t, (major): 17.2 min, tr (minor): 20.8.0 min; 65% ee; 'H-NMR (300 MHz, CDCl3): 6 = 1.48 (s, 9H), 3.94 (s, 3 H), 4.72-4.78 (m, 1H), 4.84 (dd, J = 7.5 Hz, J = 7.2 Hz, 1H), 5.63 (br s, 1 H), 5.76 (br s, 1H), 6.95-7.00 (m, 2H), 7.28-7.38 (m, 2H); l3C-NMR (300 MHz, CDC13): 6 = 180 28.3, 51.0, 55.5, 77.9, 80.3, 111.0, 121.19, 129.1, 130.9, 154.8, 156.8 ppm; MS (FAB+): m/z (%): 297 [M+H]+ (5), 252 (100), 140 (100); IR (neat): v: 3433, 2358, 2105, 1646 cm' I; mp. 138-140°C. tert-Butyl (R)-2-nitro-1-(4-methylphenyl)ethylcarbamate 43h. NHBoc N02 H3C (H)-43h According to the typical procedure, 219 mg imine were used (1 mmol) the product 43h was obtained in 45% yield (134 mg) as a white solid. HPLC analysis (Chiralpak AD, hexane/2-PrOH 98/2, flow rate = 1.0 mL/min, 1 = 210 nm): tr (minor)=50.5, tr (minor): 59.0 min; [a]25D = -26.0 (86% ee, c = 1.00, CHC13); 1H-NMR (300 MHz, CDC13): 6 =1.48 '(s, 9H), 2.37 (s, 3H), 4.70 (dd, J = 12.4, 5.7 Hz, 1H), 4.83 (s, 1H),5.40 (br s, 1 H), 5.55 (m, 1H), 7.22 (s, 4 H); 13C-NMR (300 MHz, CDC13): 6 = 21.1, 28.3, 52.7, 79.0, 80.5, 126.3, 129.8, 133.1, 138.5, 154.9 ppm. The data agreed with the previously reported literature”. tert-Butyl (R)-2-nitro-l-(l-naphthyl)ethylcarbamate 431. NHBoc l N02 (Hi-431 According to the typical procedure 255 mg imine were used (1 mmol) and the product 431 was obtained in 65% yield (165 mg, 0.65 mmol) as a white solid. HPLC analysis (Chiralpak OJ-H, hexane/2-PrOH 95/5, flow rate = 1.0 mL/min, 1 = 210 nm): tr (minor) = 35.9, tr (major) = 50.7 min; [a]25D = -8.1 (85% ee, c = 0.5, CHC13); lH-NMR 181 (300 MHz, CDC13): 6 = 1.43 (s, 9H), 4.88 (br s, 2 H), 5.25 (m, 1H), 6.27 (m, 1H), 7.45 (m, 2H), 7.54 (t, J = 7.5 Hz, 1 H), 7.61 (t,J = 7.3 Hz, 1H), 7.84 (dd, J = 5.7, 3.5 Hz,1H), 7.90 (d, J = 8.2 Hz, 1 H), 8.12 (d, J = 8.2 Hz, 1H); l3C-NMR (300 MHz, CDC13): 6 = 28.3, 49.3, 78.3, 80.7, 122.2, 123.3, 125.2, 126.3, 127.3, 129.3, 129.5, 130.3, 132.6, 134.1, 154.7 ppm; The rest of the data agreed with the previously reported literature”. tert-Butyl (R)-2-nitro-1-(3-pyridy1)ethylcarbamate 43g. NHBoc \ N02 N/ (FD-439 According to the typical procedure, 206 mg imine were used (1 mmol) and the product 43g was obtained in 63% yield (168 mg, 0.63 mmol) as a white solid. HPLC analysis (Chiralpak AD, hexane/2-PrOH 90/ 10, flow rate = 1.0 mL/min, l = 210 nm): t, (major) = 16.7, tr (minor) = 17.9 min; [a]25D = -335 (81% ee, c = 1.00, acetone); 'H—NMR (300 MHz, CDC13): 6 = 1.42 (s, 9H), 4.76 (d, J=8.9 Hz, 1H), 4.90 (s, 1 H), 5.40 (s, 2H), 7.32 (dd, J = 7.5, 4.7 Hz, 1 H), 7.65 (d, J = 7.9 Hz, 2H), 8.62 (d, J = 21.4 Hz, 2H); 13C- NMR (300 MHz, CDC13): 6 = 28.2, 50.7, 78.3, 81.0, 123.8, 133.1, 134.2, 148.1, 149.9. 154.8 ppm; The data agreed with the previously reported literature” except for the sign of the optical rotation (see discussion on data for 42). 182 tert-Butyl (1R,ZS)-2-nitro-l-phenylpropylcarbamate 43j. NHBoc N02 (1R, 23)-43] According to the typical procedure, 205 mg imine were used (1 mmol) and the product 43j was obtained in 59% yield (165 mg, 0.59 mmol) as a white solid and a 77/23 mixture of diastereomers by lH-NMR. The major diastereomer was determined to have 70% ee by chiral HPLC analysis and identification of enantiomeric pairs by UV absorption spectrum (Chiralpak AD, hexane/iPrOH 92/8 0.8 mL/min, 1 = 210 nm): Syn isomer: tr (minor) = 13.8 min, tr (major) 15.9 min; anti isomer: t, (major) = 17.8 min, tr (minor) = 21.8 min;); 1H-NMR (300 MHz, CDC13): 6 = 1.40 (s, 9H), 1.50 (d, J = 6.7 Hz, 3H), 4.91 (br s, 1 H), 5.18 (dd, J = 8.8, 5.8 Hz, 1 H), 5.32 (br s, 1H), 7.18—7.24 (m, 2H), 7.37-7.34 (m, 3 H); ”C-NMR (300 MHz, CDCl,): 6 = 154.9, 136.7, 129.0, 128.7, 128.5, 126.8, 126.4, 85.8, 80.0, 57.7, 28.3, 15.2 ppm. The data agreed with the previously reported literature”. tert-Butyl (1R,28)-2-nitro-l-phenylbutylcarbamate 43k. NHBoc No2 (1 R, 25)-43k According to the typical procedure, 205 mg imine were used (1 mmol) the product 431 was obtained as white solid in 63% yield (185 mg, 0.62 mmol) and a 80/20 mixture of diastereomers by HPLC analysis (enantiomeric pairs identified by comparison with a 183 racemic sample). The cc of major diastereomer 7k was determined to be 80% by chiral HPLC analysis (Chiralpak OJ-H, hexane/2-PrOH 97/3, 0.8 mL/min, 1 = 210 nm). Syn isomer: t, (major) = 28.1 min, tr (minor) = 50.1 min; 1H-NMR (300 MHz, CDC13): 6 = 0.96—1.03 (t, J = 3.5 Hz, 3H), 1.43 (s, 9 H), 1.84—1.92 (m, 2H), 5.10—5.14 (m, 1 H), 4.74 (br s, 1H), 5.14—5.20 (br s, 1H), 7.22-7.26 (m, 2 H), 7.30—7.40 (m, 3H); l3C-NMR (300 MHz, CDC13): 6 = 10.4, 24.8, 28.2, 56.8, 80.0, 93.0, 126.9, 128.7, 129.0, 136.7, 154.9 ppm. The data agreed with the previously reported literature”. General procedure for the synthesis of aminosulfones 62 and imines 41. NHBoc JO PhSOzNa HN ' Boc K2C03, Na2804 N , Boc ; > I A' MeOH/1420 A' 302% THF reflux Ar HCOOH 62 41 The appropriate aldehyde (1.2 eq, 20.48 mmol) was added to a mixture of 1 eq (2g, 17.07 mmol) of NHBoc and 2.0 eq. (5.6 g, 34.1 mmol) of PhSOzNa in 50 ml of water and MeOH (2:1 v/v) and stirred at rt for 3 days, after which time the resultant white suspension was filtered, washed with water and diethyl ether and then triturated with diethyl ether overnight. The product 62 was dried under vacuum. Subsequently, 1 eq. of 62 (2 mmoles) was refluxed in THF for an overnight period in the presence of 1.60 g (12 mmol) K2C03 and ~2g Na2S04 (drying agent). The resulting mixture was filtered through a clean white frit funnel and the volatiles were evaporated. The resulting imine was transferred carefully under nitrogen to the pump where it was dried under high vacuum. Cautious handling of the procedure is required for preventing decomposition and hydrolysis of the imines, which were used in the aza-Henry reaction 184 without further purification. Imines 41a, e-f and h-j are known in the literature and their preparation and data has been reported before.28 N-(ten-butoxycarbonyl)-a-(phenylsulfonyl)benzylamine 62a. NHBoc SOgPh 62a Synthesized using (2g, 17.07 mmol) of NHBoc to afford 5.89 g (17 mmol, 100%) of 623 as a white solid. 'H-NMR (300 MHz, CDC13): 6 = 1.24 (9H, s), 5.72 (1H, br d, J = 10.3 Hz), 7.49-7.38 (5H, m), 5.90 (1H, br (1, J = 11.0 Hz), 7.52 (2H, m), 7.62-7.60 (1H, m), 7.89 (2H, d, J = 7.3 Hz). N-(tert-butoxycarbonyl)-a-(phenylsulfonyl)-4—chlorobenzylamine 62b. NHBoc SOzPh Cl 621) Synthesized using 2g (17.07 mmol) of NHBoc to afford 1.48 g (3.91 mmol, 23%) of 62b as a white solid. lH-NMR (300 MHz, CDC13): 6 = 1.25 (9H, s), 5.63 (1H, br (1, J = 10.1 Hz), 5.86 (1H, br d, J = 10.1 Hz), 7.36 (4H, s), 7.52 (2H, t, J: 7.8 Hz), 7.63 (1H, t,J = 7.5 Hz), 7.88 (2H, d, J = 8.4 Hz). 185 N-(tert-butoxycarbonyl)-a-(phenylsulfonyl)-3-chlorobenzylamine 62c. NHBoc SOzPh Cl 62c Synthesized using 2g (17.07 mmol) of NHBoc to afford 4.98 g (13.9 mmol, 77%) of 62c as a white solid. 1H-NMR (300 MHz, CDCl3): 6 21.23 (9H, s), 5.69 (1H, br (1, J = 10.5 Hz), 5.87 (1H, br (1, J = 10.5 Hz),7.32-7.41 (4H, m), 7.54 (2H, t, J = 7.5 Hz), 7.62 (1H, t, J: 7.2 Hz), 7.90 (2H, d, J: 7.2 Hz). N -(ten-butoxycarbonyl)-a-(phenylsulfonyl)-2-chlorobenzylamine 62d. NHBoc SOzPh Cl 62d Synthesized using 2g (17.07 mmol) of NHBoc to afford 4.66 g (12.24 mmol, 77%) of 62d as a white solid. 'H-NMR (300 MHz, CDC13): 6 = 1.32 (9H, s), 5.85 (1H, br (1, J = 10.2 Hz), 6.65 (1H, br (1, J = 10.2 Hz), 7.37-7.48 (3H, m), 7.55-7.58 (3H, m), 7.69 (1H, t, J = 6.9 Hz), 7.97 (2H, d,J = 7.5 Hz) N-(ten-butoxycarbonyl)-a-(phenylsulfonyl)-4-bromobenzylamine 62e. NHBOC SOzPh Br 62e 186 Synthesized using 2g (17.07 mmol) of NHBoc to afford 2.24 g (5.27 mmol, 31%) of 62e as a white solid. 1H-NMR (300 MHz, CDC13): 6 = 1.25 (9H, s), 5.70 (1H, br d, J = 10.7 Hz), 5.89 (1H, br d, J = 10.7 Hz), 7.28 (2H, d, J = 8.2 Hz), 7.58-7.51 (4H, m), 7.72- 7.62 (1H, m), 7.91 (2H, d,J = 7.6 Hz). N-(tert-butoxycarbonyl)-a-(4-methylphenylsulfonyl)-4-methoxybenzylamine 62f. NHBoc SOzPh 62f Synthesized using 2g (17.07 mmol) of NHBoc to afford 5.83 g (14.28 mmol, 84%) of 62f as a white solid. 1H-NMR (300 MHz, CDCl,): 6 = 1.25 (9H, s), 2.41 (3H, s), 3.81 (3H, s), 5.79 (1H, br d, J = 10.6 Hz), 5.85 (1H, br d, J = 11.0 Hz), 6.92 (2H, d, J = 8.8 Hz), 7.31 (2H, d, J: 8.1 Hz), 7.36 (2H, d, J: 8.8 Hz), 7.78 (2H, d, J: 7.7 Hz). N-(tert-butoxycarbonyl)-or-(4-methylphenylsulfonyl)-2-methoxybenzylamine 62g. NHBoc SOgPh O/ 629 Synthesized using 2g (17.07 mmol) of NHBoc to afford 5.76 g (15.3 mmol, 90%) of 62g as a white solid. 1H-NMR (300 MHz, CDC13): 6 = 1.28 (9H, s), 2.37 (3H, s), 3.70 (3H, s), 6.16-6.27 (2H, m), 6.84 (1H, d, J = 8.1 Hz), 6.96 (1H, t, J = 7.2 Hz), 7.22-7.35 (4H, m), 7.68 (1H, d, J = 8.1 Hz); l3C-NMR (300 MHz, CDCl,): 6 = 21.6, 28.1, 55.9, 71.1, 80.8, 111.5, 118.9, 120.9, 129.4, 129.5, 130.3, 131.1, 134.6, 144.6, 154.5, 157.9; MS 187 (FAB*): m/z (%): 307 (10), 236 (100), 180 (100), 136 (90); IR (CHC13): v = 3344, 2978, 2358,1704, 1493, 1141; mp 160-162. N-(tert-butoxycarbonylya-(phenylsulfonyl)-4-methylbenzylamine 62i. NHBoc /©/‘\802Ph 621 Synthesized using 2g (17.07 mmol) of NHBoc to afford 4.54 g (12.58 mmol, 74%) of 62i as a white solid. 1H-NMR (300 MHz, CDC13): 6 = 1.28 (9H, s), 2.40 (3H, s), 5.77 (1H, d, J = 10.6 Hz), 5.90 (1H, br d, J =10.6 Hz), 7.27 (2H, d,J = 8.0 Hz), 7.34 (2H, d, J = 7.7 Hz), 7.53 (2H, t,J = 7.6 Hz), 7.64 (1H, t,J = 7.0 Hz), 7.93 (2H, d, J = Hz). N-(ten-butoxycarbonyl)-a-(phenylsu1fonyl)-C-naphthalen-l-yl-methylamine 62k. 0 NHBoc O SOgPh 62k Synthesized using 2g (17.07 mmol) of NHBoc to afford 3.98 g (10.03 mmol, 59%) of 62k as a white solid. 'H-NMR (300 MHz, CDC13): 6 = 1.27 (9H, s), 5.98 (1H, br (1, J = 10.6 Hz), 6.88 (1H, br (1, J = 10.6 Hz), 7.66-7.48 (6H, m), 7.79 (1H, d, J = 7.0 Hz), 7.88 (1H, d,J: 8.1 Hz), 7.94 (1H, d, J: 8.1Hz,), 7.99 (2H, d,J: 7.7 Hz), 8.14 (1H, d, J: 8.4 Hz). N-(tert-butoxycarbonyl)-a-(phenylsulfonyl)-C-pyridin-3-yl-methylamine 62h. 188 NHBoc / N 62h Synthesized using 2g (17.07 mmol) of NHBoc to afford 4.02 g (11.56 mmol, 68%) of 62h as a white solid. 'H-NMR (300 MHz, CDCl,): 6 = 1.25 (9H, s), 5.97 (2H, br), 7.39- 7.35 (1H, m), 7.57 (2H, t, J = 7.5 Hz), 7.68 (1H, t, J = 7.0 Hz), 7.86-7.83 (1H, m), 7.92 (2H, d, J = 8.3 Hz), 8.70-8.63 (2H, m). Benzaldehyde N-(tert-butoxycarbonyl)imine 4la. N,Boc d... Synthesized from 5.89 g (17 mmol) of 62a to afford 4.40 g (16.66 mmol, 98%) of 41a as a colorless liquid. 'H-NMR (300 MHz, CDC13): 6 = 1.59 (9H, s), 7.49-7.43 (2H, m), 7.57-7.54 (1H, m), 7.93-7.90 (2H, m), 8.88 (1H, s). p-Chlorobenzaldehyde N-(tert-butoxycarbonyl)imine 41b. .80c .04" 42b Synthesized from 1.48 g (3 .91 mmol) of 62b to afford 0.91 g (3.83 mmol, 98%) of 41a as a colorless liquid. IH-NMR (300 MHz, CDC13): 6 = 1.57 (9H, s), 7.42 (2H, d, J = 8.4 Hz), 7.85 (2H, d, J: 8.4 Hz), 8.81 (1H, s). 189 m-Chlorobenzaldehyde N-(tert-butoxycarbonyl)imine 41c. N,Boc CI 41c Synthesized using 4.98 g (13.9 mmol) of 62c to afford 3.32 g (13.76, 99%) 41c. as a clear oil. 'H-NMR (300 MHz, CDC13): 6 = 1.56 (9H, s) 7.39 (1H, t, J = 7.5 Hz), 7.48- 7.52 (1 H, m), 7.73 (1H, dd, J=1.2 Hz, J = 7.5 Hz), 8.77 (1H, s), 7.93 (1H, s). o-Chlorobenzaldehyde N-(ted-butcxycarbonylfimine 41d. Synthesized using 4.66 g (12.24 mmol) of 62d to afford 2.86 g (11.99 mmol, 99%) 41d as a clear oil. 1H-NMR (300 MHz, CDC13): 6 = 1.63 (9H, s), 7.37 (1H, t, J = 6.0 Hz), 7.47-7.50 (2H, m), 8.22 (1H, d, J :60 Hz), 9.31 (1H, s). p-Bromobenzaldehyde N-(tert-butoxycarbonyl)imine 41c. 190 Synthesized using 2.24 g (5 .27 mmol) of 62e to afford 0.95 g (3.37 mmol, 64%) 41c as a white solid. 1H-NMR (300 MHz, CDC13): 6 = 1.57 (9H, s). 7.60 (2H, d, J = 8.4 Hz), 7.77 (2H, d, J: 8.4 Hz), 8.81 (1H, s). p-Methoxybenzaldehyde N-(tert-butoxycarbonyl)imine 41f. Synthesized using 5.83 g (14.28 mmol) of 62f to afford 2.58 g (10.99 mmol, 77%) 41f as a clear oil. 1H-NMR (300 MHz, CDC13): 6 = 1.56 (9H, s), 3.86 (3H, s), 6.96 (2H, d, J = 8.8 Hz), 7.88 (2H, d,J: 8.8 Hz), 8.88 (1H, s). o-Methoxybenzaldehyde N-(tert-butoxycarbonyl)imine 41g. Synthesized using 5.76 g (15.3 mmol) of 62g to afford 3.41 g (14.53 mmol, 95%) 41g as a clear oil. 1H-NMR (300 MHz, CDC13): 6 = 1.56 (9H, s), 3.88 (3H, s), 6.90-7.00 (2H, m), 7.49 (1H, t, J = 9.0 Hz), 8.09 (1H, d, J = 7.8 Hz), 9.35 (1H, 5); ”GM (300 MHz, CDC13): 6 = 27.9, 55.6, 81.9, 111.27, 120.77, 122.6, 128.3, 135.2, 160.9, 165.8, 163.2; MS (FAB*): m/z (%): 236 [M+H]+ (60), 180 (100), 136 (95); IR (CHC13): v=2978, 1711,1600,1237,1153. 191 p-Tolualdehyde N-(tert-butoxycarbonyl)imine 4li. N,Boc fig 411 Synthesized using 4.54 g (12.58 mmol) of 62i to afford 2.67 g (12.20 mmol, 97%) 4li as a white solid. 1H-NMR (300 MHz, CDC13): 6 = 1.58 (9H, s), 2.41 (3H, s), 7.26 (2H, d,J: 7.7 Hz), 7.81(2H, d,J = 8.1 Hz), 8.87 (1H, s). l-Naphthaldehyde N-(ten-butoxycarbonymmine 41k. 0 IN’BOC 41k Synthesized using 3.98 g (10.03 mmol) of 62k to afford 2.55 g (10.00 mmol, 99%) 41k as a white solid. 'H-NMR (300 MHz, CDC13): 6 = 1.61 (9H, s), 7.60-7.54 (2H, m), 7.68-7.63 (1H, m), 7.90 (1H, d, J = 7.7 Hz), 8.05 (1H, d, J = 8.4 Hz), 8.17 (1H, d, J = 7.0 Hz), 8.92 (1H, d, J: 8.4 Hz), 9.53 (1H, s). 3-pyridinecarboxaldehyde N-(tert-butoxycarbonyl)imine 41h. 192 Synthesized using 4.02 g (11.56 mmol) of 62h to afford 2.32 g 41h (11.3 mmol, 98%) as a clear oil. 1H-NMR (300 MHz, CDCl3): 6 = 1.58 (9H, s), 7.40 (1H, dd, J = 4.8, 8.1 Hz), 8.25-8.28 (1H, dt, J = 1.8, 8.1 Hz), 8.76 (1H, dd, J = 1.83, 4.8 Hz), 8.87 (1H, s), 9.00 (1H,d,J= 2.2 Hz). 193 Experimental Procedures for Chapter 5 General procedure for the non-asymmetric addition of l-nitropropane to nitrostyrene using catalyst 47 (Table 28). 2% FC. ; ii ixi ; .CF 3 \n/ 3 S R/VNOZ CF3 CF3 N (022M) 2% EtaN 02 > R N02 + /\/N02 toluene, rt 20 eq 803-1 A flame-dried round bottom flask was charged with 1 mmol of the appropriate nitrostyrene and 10 mg (0.02 mmol) of catalyst 47 and the solids were dissolved in 3 mL of toluene. 1-Nitropropane (1.6 ml, 20 mmol) was added followed by 3uL of EN. The mixture was stirred at r.t. for the time indicated in the Table. The volatiles were evaporated and the crude product was purified with silica gel column chromatography to afford the racemate products 80a-i as separated diastereomers. The syn/anti ratio was measured by the weight of the isolated products except in the case of 80a, 80b where it was determined by lH-NMR analysis. 1. 94 General procedure for the asymmetric addition of l-nitropropane to nitrostyrene using catalyst (R)-82 (Table 39). NO R “~02 + /\/N02 2% mol. 82 2 > N02 (0.2M) 30 eq benzene, rt R BOa-l A flame-dried round bottom flask was charged with 1 mmol of the appropriate nitrostyrene and 13.5 mg (0.02 mmol) of catalyst 82 and the solids were dissolved in 2.5 mL of benzene. 1-Nitropropane (2.5 mL, 30 mmol) was added and the mixture was stirred at r.t. for the time indicated in the Table 39. The volatiles were evaporated and the crude product was purified with silica gel column chromatography to afford the products 80a-i as separated diastereomers. The syn/anti ratio was measured by the weight of the isolated products except in the case of 80a, 80b, 80c, 80d where it was determined by 1H- NMR analysis. 2-Phenyl-1,3-dinitropentane 803 N02 No2 (2H, 3F0-80a Prepared following the general procedures described above using 150 mg (1.00 mmol) nitrostyrene. Purified with silica gel column using 30% EtOAc/hexanes to afford 144 mg (60%, 0.60 mmol) of syn adduct as a clear oil and 47 mg (0.20 mmol, 20%) anti adduct from the asymmetric addition (Table 39). Syn adduct: 'H-NMR (300MHz), CDCl3 6 =1.00 (t, 3H, J = 7.1 Hz), 1.90-1.82 (m, 1H), 2.02-1.94 (m, 1H), 4.02 (q, 1H, J = 6.7 Hz), 4.78-4.71 (m, 2H), 4.86 (dd, 1H, J = 6.3, 13.9 Hz), 7.14-7.11 (m, 2H), 7.33-7.29 (m, 195 3H); l3C-NMR (300MHz), CDC], 6 = 10.3, 24.4, 46.6, 76.4, 91.2, 127.9, 128.1, 129.3, 133.9 ppm; R, = 0.30 (30% EtOAc/hexanes); HPLC analysis: Chiralpac OJ—H, hexane/2- PrOH 80/20, 1.5 mL/min, 95% ee, (1”,, = +7.2 (c=1.0, dichloromethane), tr = 31.0 min (major), tr = 42.0 min (minor). The data agreed with the reported literature“. The non-asymmetric addition (Table 28) afforded 178 mg (75%) of syn-80a adduct and 28 mg (0.12 mmol, 12%) anti adduct. 2-(4-Methoxyphenyl)-1,3-dinitropentane 80b NO2 N02 MeO (2R, 3H)-80b Prepared following the general procedures described above using 179 mg (1.00 mmol) of the nitrostyrene. Purified with silica gel column using 20% EtOAc/hexanes to afford 126 mg (47%, 0.47 mmol) of syn adduct as a clear yellow oil and 34 mg (13%, 0.13 mmol) anti adduct from the asymmetric addition (Table 39). Syn adduct: lH-NMR (300MHz), CDC], 6 = 0.99 (t, 3H, J = 7.1 Hz), 187-1.1.80 (m, 1H), 2.02-1.93 (m, 1H), 3.75 (s, 3H), 3.95 (q, 1H, J = 6.7 Hz), 4.72-4.71 (m, 2H), 4.83 (dd, 1H, J: 6.3, 13.9 Hz), 7.04 (d, J = 8.7 Hz, 2H), 7.84 (d, J = 8.7 Hz, 2H); l3C-NMR (300MHz), CDC], 6 =10.5, 24.6, 46.2, 55.5, 76.8, 91.5, 114.9, 125.8, 129.3, 160.2; R, = 0.08 (20% EtOAc/hexanes); HPLC analysis: Chiralpac OD-H, hexane/2-PrOH 70/30, 0.5 mL/min, 94% ee, (1”,, = +220 (c=1.0, dichloromethane), tr = 15.6 min (major), tr = 31.1 min (minor). The data agreed with the reported literature“. 196 The non-asymmetric addition (Table 28) afforded 197 mg (0.73 mmol, 73%) of syn-80a adduct and 24 mg (0.09 mmol, 9%) of the anti adduct. 2-(2-Methoxyphenyl)-1,3-dinitropentane 80c No2 NO2 OMe (2R, 3H)-80c Prepared following the general procedures described above using 179 mg (1.00 mmol) of the nitrostyrene. Purified with silica gel column using 20% EtOAc/hexanes to afford 193 mg (72%, 0.72 mmol) of syn adduct as a clear oil and 58 mg (0.22 mmol, 22%) of the anti adduct from the asymmetric addition (Table 39). Syn adduct: lH-NMR (300MHz), CDC1,6 = 0.98 (t, 3H,J= 7.1 Hz), 1.93-1.82 (m, 1H), 2.03-1.94 (m, 1H), 3.85 (s, 3H), 4.24 (q, 1H, J = 7.5 Hz), 4.84 (dd, 2H, J = 5.1, 7.5 Hz), 5.00 (m, 1H), 6.88-6.86 (m, 2H), 7.04 (dd, J = 1.6, 7.5 Hz, 1H), 7.28-7.25 (m, 1H). l3C-NMR (300MHz), CDC], 6 = 10.6, 24.8, 43.9, 55.7, 75.6, 90.2, 111.6, 121.4, 122.4, 130.4, 130.5, 157.5 ppm; R, = 0.20 (20% EtOAc/hexanes); HPLC analysis: Chiralpac OD-H, hexane/Z-PrOH 80/20, 0.5 mL/min, 94% ee, a251, = +8.0 (c=1.0, dichloromethane), tr = 11.9 min (major), tr = 23.3 min (minor). The data agreed with the reported literature“. The non-asymmetric addition (Table 28) afforded 178 mg (0.66 mmol, 66%) of syn-80c adduct and 45 mg (0.17 mmol, 17%) of the anti adduct. 197 4-Methyl-1,3-dinitropentane 80d N02 N02 H3C (29, 3H)-80d Prepared following the general procedure described above using 163 mg (1.00 mmol) of the nitrostyrene. Purified with silica gel column using 20% EtOAc/hexanes to afford 146 mg (58%, 0.58 mmol) of syn adduct as a clear oil and 35 mg (0.14 mmol, 14%) from the asymmetric addition (Table 39). Syn adduct: 'H-NMR (300MHz), CDC], 6 = 0.98 (t, J = 7.2 Hz, 3H), 1.80-1.84 (m, 1H), 2.02-1.96 (m, 1H), 2.29 (s, 3H), 3.97 (q, J = 8.1 Hz, 1H), 4.75-4.68 (m, 2H), 4.80 (dd, J = 13.5, 6.3 Hz, 1H), 7.10 (d, J = 8.1 Hz, 2H), 7.12 (d, J = 8.1 Hz, 2H); l3C-NMR (300MHz), CDC], 6 = 10.3, 21.1, 24.4, 46.2, 76.5, 91.2, 127.8, 129.9, 130.8, 138.9 ppm; R, = 0.17 (20% EtOAc/hexanes); MS (FAB'): m/z (%): 251 [M-H]' (60), 239 (25), 204 (60); IR (neat) 3028, 2978, 2926, 2883, 1552, 1510, 1460, 1435, 1377, 1318, 1122 cm"; HPLC analysis: Chiralpac OD-H, hexane/2-PrOH 70/30, 0.5 mL/min, 94% ee, 0”,, = +11.0 (c=1.0, dichloromethane), tr = 14.8 min (major), tr = 45.7 min (minor). The non-asymmetric addition (Table 28) afforded 142 mg (0.56 mmol, 56%) of syn-80d adduct and 40 mg (0.16 mmol, 16%) of the anti adduct. 198 2-(2-Chlorophenyl)-1,3-dinitropentane 80c no2 N02 Cl (23, 3H)-BOe Prepared following the general procedure described above using 183 mg (1.00 mmol) of the nitrostyrene. Purified with silica gel column using 15% EtOAc/hexanes to afford 175 mg (0.62 mmol, 62%) of syn adduct as a clear oil and 60 mg (0.22 mmol, 22%) of the anti adduct from the asymmetric addition (Table 39). Syn adduct: 'H-NMR (300MHz), CDC], 6 = 1.06 (t, J = 7.2 Hz, 3H), 2.08-2.04 (m, 2H), 4.70-4.67 (m, 1H), 4.90 (dd, J = 5.9, 8.3 Hz, 2H), 5.02-5.07 (m, 1H), 7.17-7.20 (m, 1H), 7.29-7.45 (m, 2H), 7.49- 7.46 (m, 1H); l3C-NMR (300MHz), CDC], 6 = 10.6, 24.3, 43.0, 75.4, 90.2, 127.9, 128.8, 130.4, 130.9, 132.0, 134.5; R, = 0.22 (15% EtOAc/hexanes); HPLC analysis: Chiralpac OJ-H, hexane/2-PrOH 70/30, 0.5 mL/min, 92% ee, (1”,, = +6.7 (c=1.0, dichloromethane), tr = 48.4 min (major), tr = 74.9 min (minor). The data agreed with the reported literature48 except for the sign of the optical rotation. The non-asymmetric addition (Table 28) afforded 185 mg (68%) of syn-80c adduct and 32 mg (0.12 mmol, 12%) of the anti adduct. Data for anti-80c: lH-NMR (300MHz), CDC], 6 = 0.93 (t, J = 7.5 Hz, 3H), 1.61- 1.69 (m, 1H), 1.95-2.00 (m, 1H), 4.55-4.62 (m, 1H), 4.69 (dd, J = 3.9 Hz, 13.5 Hz, 1H), 4.80-5.00 (m, 1H), 7.14-7.44 (m, 4H); l3C-NMR (300MHz), CDC], 6 = 10.2, 25.4, 42.6, 75.0, 90.76, 103.9, 127.9, 130.2, 130.9, 131.9; one C not located; a251, = +8.0 (c=0.8, dichloromethane). 199 2-(4-Chlorophenyl)-1,3-dinitropentane 80f no2 N02 Cl (2 H, 33)-80f Prepared following the general procedure described above using 183 mg (1.00 mmol) of the nitrostyrene. Purified with silica gel column using 20% EtOAc/hexanes to afford 168 mg (0.62 mmol, 62%) of syn adduct as a clear oil and 35 mg (0.13 mmol, 13%) anti adduct from the asymmetric addition (Table 39). Syn adduct: lH-NMR (300MHz), CDC],6 = 0.99 (t, J = 7.2 Hz, 3H), 1.80-1.90 (m, 1H), 1.96-1.99 (m, 1H), 4.00 (q, J: 7.8 Hz, 1H), 4.75-4.67 (m, 2H), 4.83 (dd, J = 13.5, 6.0 Hz, 1H), 7.08 (dd, J = 6.6, 2.4 Hz, 2H), 7.31 (dd, J = 6.6, 2.4 Hz, 2H). l3C-NMR (300MHz), CDC],6 = 10.2, 24.5, 45.9, 76.2, 90.9, 129.3, 129.5, 132.3, 135.]; MS (FAB'): m/z (%): 273 [M-H]' (27, 37Cl), 271 [M-H]' (80, 35Cl) 224 (100); IR (neat) 3406 (H20), 2980, 1554, 1493, 1460, 1435, 1377, 1095, 1014 cm"; HPLC analysis: Chiralpac OD-H, hexane/Z-PrOH 70/30, 0.5 mL/min, 92% cc, 0125,, = +10.9 (c=1.0, dichloromethane), tr = 13.7 min (major), tr = 24.8 min (minor). For the anti adduct (x250: +23.2 (c=1.0, dichloromethane). The non-asymmetric addition (Table 28) afforded 157 mg (58%) of syn-80a adduct and 40 mg (0.15 mmol, 15%) of the anti adduct. 2-(4-Bromophenyl)-1,3-dinitropentane 80g. NO2 N02 Br (2R, 3H)-809 200 Prepared following the general procedure described above using 228 mg (1.00 mmol) of the nitrostyrene. Purified with silica gel column using 20% EtOAc/hexanes to afford 207 mg (0.65 mmol, 65%) of syn adduct as a clear oil and 31 mg (0.10 mmol, 10%) of the anti adduct from the asymmetric addition (Table 28). Syn adduct: lH-NMR (300MHz), CDC],6 = 1.00 (t, J = 7.2 Hz, 3H), 1.88-1.81 (m, 1H), 2.03-1.96 (m, 1H), 3.99 (q, 6 Hz, 1H), 4.67-4.76 (m, 2H), 4.83 (dd, J = 19.5, 6 Hz, 1H), 7.02 (d, J = 9 Hz, 2H), 7.46 (d, J = 9 Hz, 2H). I3C-NMR (300MHz), CDC], 6 = 10.2, 24.4, 46.0, 76.0, 90.8, 123.3, 129.6, 132.5, 132.8 ppm; R, = 0.11 (20% EtOAc/hexanes); IR (neat) 3451(H20), 2112, 1653, 1558, 1375; MS (FAB'): m/z (%): 317 [M-H]' (98, S'Br), 315 [M-H]' (97, 7S’Br), 268 (100); HPLC analysis: Chiralpac OD-H, hexane/Z-PrOH 70/30, 0.5 mL/min, tT = 14.8 min (major), tr = 31.8 min (minor), 93% ee, 0,st = +10.2 (c=1.0, dichloromethane). The non-asymmetric addition (Table 28) afforded 180 mg (0.57 mmol, 57%) of syn-80a adduct and 28 mg (0.09 mmol, 9%) of the anti adduct. 2-(3-Bromophenyl)-l,3-dinitropentane 80h. N02 N02 Br (21?, 3Fi)-80h Prepared following the general procedure described above using 228 mg (1.00 mmol) of the nitrostyrene. Purified with silica gel column using 20% EtOAc/hexanes to afford 197 mg (0.62 mmol, 62%) of syn adduct as a clear yellow oil and 22 mg (0.07 mmol, 7%) of the anti adduct from the asymmetric addition (Table 39). Syn adduct: 1H- 201 NMR (300MHz), CDC], 6 = 1.00 (t, J = 7.5 Hz, 3H), 1.88-1.81 (m, 1H), 2.03-1.94 (m, 1H), 3.97 (q, 6 Hz, 1H), 4.77-4.70 (m, 2H), 4.83 (dd, J = 13.5, 6 Hz, 1H), 7.07 (d, J = 8.1 Hz, 1H), 7.21 (t, J = 7.8 Hz, 1H), 7.30 (t, 1.8 Hz, 1H), 7.47 (dq, J = 8.1 Hz, J = 1.2 Hz, 1H). l3C-NMR (300MHz), CDC], 6 = 10.2, 24.5, 46.1, 76.0, 90.9, 123.2, 126.5, 130.9, 131.2, 132.3, 136.3; MS (FAB'): m/z (%): 317 [M-H]' (98, 8'Br), 315 [M-H]‘ (97, 7S’Br), 268 (100); IR (neat) 3451 (H20, 2112, 1653, 1558, 1375, 1010; HPLC analysis: Chiralpac OD—H, hexane/2-PrOH 70/30, 0.5 mL/min, 94% ee, 0251) = +4.3 (c=1 .0, dichloromethane), tr = 14.5 min (major), tr = 69.6 min (minor). The non-asymmetric addition (Table 28) afforded 217 mg (0.69 mmol, 69%) of syn-80h adduct and 19 mg (0.06 mmol, 6%) of the anti adduct. 2-(2-Bromophenyl)1,3-dinitropentane 80i NO2 N02 Br (21?, {Sm-80] Prepared following the general procedure described above using 228 mg (1.00 mmol) of the nitrostyrene. Purified with silica gel column using 20% EtOAc/hexanes to afford 185 mg (0.58 mmol, 58%) of syn adduct as a clear oil and 63 mg (0.20 mmol, 20%) of the anti adduct from the asymmetric addition (Table 39). Syn adduct: 'H-NMR (300MHz), CDC], 6 = 1.00 (t, J = 7.5 Hz, 3H), 1.94-2.04 (m, 2H), 4.62-4.70 (m, 1H), 4.82-4.91 (m, 2H), 4.95-5.04 (m, 1H), 7.11-7.21 (m, 2H), 7.29 (td, J = 7.5, 1.2 Hz, 1H), 7.61 (dd, J = 8.1, 1.5 Hz, 1H); l3C-NMR (300MHz), CDC], 6 = 10.4, 24.0, 44.8, 75.3, 90.0, 128.3, 130.4, 133.4, 134.]; MS (FAB'): m/z (%): 317 [M-H]' (98, S'Br), 315 [M-H]' 202 (97, 79Br), 268 (100); IR (neat) 3451(H20), 2112, 1653, 1558, 1375, 1010; HPLC analysis: Chiralpac OD—H, hexane/2-PrOH 70/30, 0.5 mllmin, t, = 11.8 min (major), tr = 38.5 min (minor), 92% ee, (1”,, = -6.8 (0:10, dichloromethane). For the anti adduct 0:25., = +6.6 (c=1 .0, dichloromethane). The non-asymmetric addition (Table 28) afforded 191 mg (60%) of syn-80i adduct and 19 mg (0.06 mmol, 6%) of the anti adduct. 2-Phenyl-1 ,3-dinitropentane 80j N02 N02 enamem Prepared following the general procedure described above, using 50 mg (0.33 mmol) of nitrostyrene. Purified with silica gel column using 20% EtOAc/hexanes to afford 63 mg (0.23 mmol, 70%) of syn adduct as a white solid and 20 mg (0.08 mmol, 8%) of the anti adduct. Syn adduct: 'H-NMR (300MHz), CDC], 6 = 0.94 (t, J = 7.2Hz, 3H), 1.34 (tq, J = 7.2 Hz. 2H), 1.74-1.68 (m, 1H),2.00—1.91 (m, 1H), 4.00 (t, J = 6.9 Hz, 1H), 4.92-4.72 (m, 3H), 7.15-7.11 (m, 2H), 7.34-7.31 (m, 3H) ppm; 13C-NMR (300MHz), CDC], 6 = 13.3, 19.1, 32.7, 46.7, 76.2, 89.3, 127.8, 129.0, 129.2, 133.7 ppm; HPLC analysis: Chiralpac OJ-H, hexane/Z-PrOH 80/20, 1.5 mllmin, tr = 26.2 min (major), tr = 31.9 min (minor), 91% cc, (1251, = --(c=1.0, dichloromethane); The data agreed with the reported literature“. 203 Thioureas 47 and 76-78 were prepared according to literature procedures“c and their spectroscopic data matched those of the literature. Thiourea 79. CF3 CF3 ©:NH2 scr‘1/