RETUZINUI‘G LIBRARY #3759351" ‘7' M 22¢ per day per itu; Place in bcuk return ta rt:.;e charge from circa1;51:n recards MULTINUCLEAR NMR STUDY ON THE HIGHLY BRIDGED ARYL SUBSTITUTED COATES CATION AND SIMILAR STUDIES ON THE CLASSICAL ARYL SUBSTITUTED 2-ADAMANTYL AND 7-NORBORNYL CATIONS By Thomas P. Clausen A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1980 ABSTRACT MULTINUCLEAR NMR STUDY ON THE HIGHLY BRIDGED ARYL SUBSTITUTED COATES CATION AND SIMILAR STUDIES ON THE CLASSICAL ARYL SUBSTITUTED 2-ADAMANTYL AND 7-NORBORNYL CATIONS By Thomas P. Clausen Following the work of Botto and Chambers, three series of aryl 'substituted carbocations were prepared and their NMR spectra taken. The first series studied was the 9-ary1-9-pentacyclo[A.3.0.0?’40?’80§’Z) nonyl cation; the only system that all in the field agree to be bridged. The results from the study of it are twofold: 1) The NMR of aryl substituted bridged ions is in sharp contrast to those of classical ions. The degree of contrast is difficult to predict but qualitative trends are readily apparent. 2) The perturbation of placing an aryl substituent on a bridged system tends to stabilize the "classical" struc- ture much more than the bridged structure. Thus weakly bridged systems may easily become classical by the place- ment of even very electron demanding aryl groups. The second system studied was the 2-aryl-2-adamantyl cation. It was found that the perturbation of placing an aryl group on the weakly- bridged 2-adamanty1 cation produces a classical cation. Hence a Thomas P. Clausen lower limit for the amount of bridging that is detectable by this probe has been established. The third system studied was the 7-aryl-7—norbornyl cation with the hope of inferring some unusual behavior for the parent cation. No such behavior was found, however, and hence, if the geometry of the 7-norborny1 cation is deformed, then the energy difference between it and its classical structure must be too small to measure by the probe used. -TABLE OF CONTENTS Page INTRODUCTION 0 O O 0 O O O O O O O O O O O O O O O O O O O O O O 1 Review of the Coates cation. . . . . . . . . . . . . . . 14 Review of the 2-adamanty1 cation . . . . . . . . . . . . 18 Review of the 7-norborny1 cation . . . . . . . . . . . . 26 RESULTS. 0 O 0 O O O O O O O O O O O O O O O O O O O O O O O O I 31 DISCUSSION 0 O O .0 O O O O O O O O O O O O O O O O O O O O O O O 41 Coates system. . . . . . . . . . . . . . . . . . . . . . 41 Z-Adamantyl SYStem o o o o o o o o o o o o o o o o o o o 49 7-N0rb0rny1 SYStem o o o o o o o o o o o o o o o o o o o 53 EXPERIMENTAL O O O O O O O O O O O O O O O O O O O I O O O I I O 56 Instrumentation. . . . . . . . . . . . . . . . . . . . . 56 Carbocation precursor formation. . . . . . . . . . . . . 56 Carbocation formation. . . . . . . . . . . . . . . . . . S7 Quench study on (83) . . . . . . . . . . . . . . . . . . 59 Measurement of the dependence of the 13C NMR of (27a) on acid concentration . . . . . . . . . . . . . . . 59 Physical and spectral data of the carbocation precursors . . . . . . . . . . . . . . . . . . . 60 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 ii LIST OF TABLES TABLE PAGE 1. NMR chemical shifts of (68) and (69) . . . . . .I. . . . . 25 2. 1H NMR chemical shifts downfield from ms of (26). . . . . 32 3. 13C NMR chemical shifts downfield from TMS of (26) . . . . 33 4. 1H NMR chemical shifts downfield from ms of (27). . . . . 34 5. 13C NMR chemical shifts downfield from TMS of (27) . . . . 35 6. Dependence of 13C NMR chemical shifts of (27a) with re- spect to acid concentration. . . . . . . . . . . . . . . . 37 7. 1H NMR chemical shifts downfield from TMS of (28) and (83) . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 8. 13C NMR chemical shifts downfield from TMS of (28) and (83) . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 9. Comparison of the 1H NMR of the ortho hydrogens between (26), (27) and (28). . . . . . . . . . . . . . . . . . . . 45 iii FIGURE 10. ll. 12. 13. 14. LIST OF FIGURES Correlation of the chemical shift of (15) with:’+. . Correlation of the chemical shift of (16) with ffil . Correlation of the chemical shift of (17) with 0+. Plot of the chemical shifts of the cationic center of (15) VS. (19). o o o o o o o o e o o o o o o o 0 Plot of the chemical shifts of the cationic Center of (19) V8. (22). o o o o o o o o o o o o o o o o o 0 Plot of the chemical shifts of the cationic center of (25) vs. (20). . . . . . . . . . . . . . . . . Calculated geometry of the 7-norbornyl cation. . . . 13C NMR chemical shifts of the cationic carbon of (26) VS. (19) o o o ., o o o o o o o o o o o o o o o o o o o 13C NMR chemical shifts of the cationic carbon of (84) VS. (19) o o o o o o o o o o o o o o o o o o o o o o o A qualitative energy diagram for the bridge flipping of (45). . . . . . . . . . . . . . . . . . . . . . . A modification of figure 10 stressing the higher energy of aryl substituted bridged cations compared to classi- cal cations. . . . . . . . . . . . . . . Molecular orbital diagram showing the interaction of aryl groups with a bridged system. . . . 13C NMR chemical shifts of the cationic carbon of (27) vs. (19) . . . . . . . . . . . . . . . 1 H NMR chemical shift of the nonequivalent geminal pro- tons in (27) plotted against each other. . iv Page 10 11 3O 42 43 46 47 48 SO 51 LIST OF FIGURES--continued FIGURE Page 13C NMR chemical shifts of the cationic carbon of (28) vs. (19). . . . . . . . . . . . . . . . . . . . . . . . 54 15. 16. Bond polarization of the endo hydrogens of (28) by through space interaction with the cationic center. . . 55 .INTRODUCTION In 1939, Wilson1 made the suggestion that the proposed intermed- iates (3) and (4) in the solvolysis of camphene hydrochloride (1) might exist as a resonance hybrid (5) rather than separate entities. In the early 1960's, structures of other cations showing similar delocaliza- tion to that of (S) were proposed, the most famous being the 2-norbornyl (6)2 and the cyclopropylcarbinyl (7)3 cations. Soon, practically every carbocation known was reevaluated in terms of delocalized sigma bonds and was given the designation of a "nonclassical cation". It was not until the late 1970's, however, when the term "nonclassical cation" was defined.48 cu on (I) . (2) (3) (4) (6) . m In the early 1960's, when H.C. Brown5 and others challenged the issue of nonclassical cations, an immense amount of extremely valuable work was undertaken to prove or disprove the existence of these ions. Possibly no other area of organic chemistry has had the time and scrutiny of so many chemists as this question concerning the carbocation geometry. Although even after a quarter of a century, many of the early questions raised are still the objects of enthusiastic debate (the 2- norbornyl and cyclopropylcarbinyl cationic structures for instance), there is no doubt that these questions have inspired a great deal of advancement in organic chemistry. At present there can be no doubt as to the possibility of nonclas- sical cations existing. The threercenter, two-electron bonds foUnd in many boranes strongly indicate that carbocations (which are isoelec- tronic with the boranes) may also enjoy the same type of bonding. Recen- tly, structures (8) - (11) have gained wide approval over their classical counter-parts.6'9 Still, however, the approval of structures (8) - (11) is more than balanced by the rejection of a great many proposed nonclassi- cal structures4b such as (12) - (14). + 9H3 (I 2) (I 3) (I 4) There remains, however, a large number of carbocations whose structures have yet to be established as classical or nonclassical. Of the many methods used to prove the structure of carbocations, NMR seems to have been the most used. The main disadvantage of NMR studies has been the lack of suitable model compounds. Thanks to the efforts of Olah and others,10 we now have a workable supply of NMR data on proven classical cations. The limited number of cations known to be bridged, however, offers us little in the way of what is "normal" for the NMR of . nonclassical cations. The major theme of this thesis is to enhance our knowledge of what is normal for classical and nonclassical cations. The method used to examine cations in this project was to correlate chemical shifts of variousaC-aryl cations with those of known classical cations as a function of electron demand. Olah has shown that a plot of the chemical shift of the cationic center of cations (15) - (l7) corre- lates in a somewhat linear fashion wither+ (see figures 1-3).11 From the large standard deviations, however, it became clear that the effect of different aryl groups on the chemical shifts was not just a function of electron demand, and that a new set of SOLVOLYSIS “ R2 PRODUCTS (67) SUBSTRATE RELATIVE RATE RT = R2” H l Rle, R2=CH3 14—21 R,= R2= CH3 38—70 25 (70) Table 1 NMR chemical shifts (PPM from TMS) of (68) and (69). Cation] CL?) '62 54,3310 C57 TCG ‘36 l38.6 92.3 5).o 26.2 46.7 (68) 'H am 2.53 l.68 (69, 'f’c 68.3 329.1 63.? 9.89.... H 2.68 2.08 Scheme IV 26 All of the evidence presented strongly suggests a bridged struc- ture for 2D 2-adamanty1 cations. However, it is prudent to reemphasize that if such bridging is in fact present, it is unsymmetrical, and hence, probably weak. For instance, the excellent correlation of 2-adamantyl tosylate with regard to the Foote-Schleyer equation38 is consistent with little or no bridging of the 2-adamantyl cation during solvolysis. 7-norbornyl cation (30) As a result of the instability of the 7-norborny1 cation, studies associated with it and its precursors have been meager. In addition to the scant amount of data concerning it, interpretations of such data have been widely divergent, leading to three structures, (30),38 (72)39’40-42 and (73),“3 proposed for this elusive cation. The arguments which support these structures will be briefly discussed below. (30) ' (72) (73) The first representation of the 7—norbornyl cation is that shown by (30); i.e., a classical cation. Correlation with the Foote-Schleyer 1c . 38 obs _ _ , ca equation (log kre — 7.00, log krel 1 = ’7-04) is the only direct evi- dence suggesting such a structure. However, from the scattering of data points about the Foote-Schleyer equation, a rate enhancement by a factor. of ten could be easily undetected implying the possibility of a small amount of anchimeric assistance in the solvolysis of 7-norborny1 tosylates. Furthermore, as stated earlier, the absence of any anchimeric assistance 27 in the transition state of such a solvolysis implies little about the nature of the intermediate. Perhaps the strongest case for (30) is the lack of sufficient evidence to prove the existence of the more controversial structures, (72) and (73). Nonetheless, evidence does exist for (72) and (73) and is given below along with some classical interpretations. Structure (72) was first postulated by Winstein and co—workers39 in 1958. Winstein's conclusions were based primarily on the products obtained from the acetolySis of 7-norbornyl brosylate (74) and its isomer, 2—bicyclo[8.2.0]heptyl brosylate (75). The mechanism proposed invoked the intermediate (72) to account for the absence of any of the endo isomer of (76); i.e., (79). Similar arguments have been forwarded for the exclu- sive formation of exo-2-norborny1 products in the solvolysis of endo and 44 Since this work, however, Brown45 has exo-2-norbornyl derivatives. refuted the need to impose nonclassical intermediates on the basis of the exo/endo ratio of product distributions. To be explicit, regardless of whether or not the intermediate is bridged, the exo face is the less hindered and hence should capture the solvent preferentially. Later, Gassman and others found additional evidence supporting Winstein's proposed intermediate.l‘0'42 Gassman found that the acetolysis of exo, exo-2,3-dideuterio-anti-tosyloxynorbornane (80) and the syn isomer, yields the corresponding acetate (82) with about 90% retention. From this, it is clear that the free 7-norborny1 cation (30) plays, at best, a minor role in the solvolysis. Two possibilities were considered or the intermediate, one being a nonclassical intermediate such as (dz-72) and the other being a tight ion pair (81) that rapidly collapses to (82). 28 CBS (75). (72) . (74) l H OAc (77) (78) 0.46 , (76) I‘llélé f1 (79) The presence of about 10% inversion in the product is most readily explained by conversion of either (dz-72) or (81) to the free classical cation (31) followed by solvent capture. An SN2 type mechanism is un- likely because the addition of acetate ion has no observable effect on 41 the percent retention of configuration found in the product. The pos- sibility of sulfur—oxygen bond breakage instead of carbon-oxygen bond clea- 18 vage of (80) to explain the amount of retention was excluded by an O- labeling experiment.46 29 H 0T5 . ' (81) (8°) (32) (dz-72) Although structure (73) could also be used to explain Gassman's observations, it was not considered. The only mention of (73) was in a paper by Dewar43 describing a MINDO/Z calculation to obtain an optimum geometry for the 7-norbornyl cation. His calculations assumed a plane of symmetry passing through atom 7 and bisecting the 2,3 and 5,6 bonds (thereby excluding the possibility of (72)). The results of his calcu- lations clearly favor an unsymmetrical cation as diagrammed on the following page. It is evident that very little can be conclusively stated concerning the exact geometry of the 7-norbornyl cation. Perhaps Brown's arguments should be applied here and the classical structure (30) accepted until more convincing evidence for (72) and (73) can be forwarded. 30 Figure 7 Calculated geometry of the 7-norbornyl cation H .7 l76.9° RESULTS The details for the generation and observation of the carbocations are given in the experimental section. The results of these observations are briefly stated below. The 1H NMR chemical shifts of thecx-aryl Coates cations (26a—e) are presented in tables 2 and 3 respectively. The spectra were unaffected from -100°C to about -500C at which time decomposition occurred. The chemical shifts of the 9-pheny1 Coates cation (26b) have been reported 8b and are included in the tables. earlier Several unsuccessful attempts were made to generate the 9-(p-anisyl) Coates cation (26f) and the 9-(3,5-dichloro-4-N,N-dimethylammoniophenyl) Coates cation (26g). In the case of the electron rich cation (26f), the complexity of the spectra observed suggested decomposition. For (26g), however, rearrangement to two species predominated even at -100°C. No attempts were made to elucidate the structures of the rearranged products. Cl + CH3 4 NH(CH3)2 + Cl 1H NMR and 13C NMR spectra of the a-aryl adamantyl cations The (27a-i) are given in tables 4 and 5 respectively. Except for (27i), all of these systems were stable up towards the boiling point of the solvent (+1OOC). 31 Amnmmv aow.~( Annmav H~¢.n)n< Anm h.wuhmvv «ow.n) Ana m.wuhmvv mom.nlu< Asmmmv mow.e(mo~.~) Aux w.ouhmuv mmm.~Iu¢ Am©.5vl ABV omc.hlm¢m.nlu< Aux m.nahmwv mnw.nl Aha n.eusuev oma.mmu< euc.uu mo .omHz Nmo.m owm.m wmn.m Am~.~v oqm.m naa.N no «0 .nm monoummmu Eoum coxmu mum mammnuaoume a“ enemaszm omo.~ nmq.m mam.m Ame.mv mmm.m oew.m 5 mu m No wmm.q omn.q mHo.m Aoo.ev moo.¢ mmooINAm movum.m c o m m UI mole e o m Ulmlm mmoo emeoummone msouw Hmu< Aoomv Acemv Aoemv mAnomv Amway mcsomaoo .Aemv mo Amze seem ammo humane amasseee mzz ma N wanmw 32 .nm mocmumwou Boom coxmu mum wfimozucoume CH mumn552m me.oe . eo.oe me.em Ha.me mmeoumflmeov-m.m AceNe ee.oe . om.em He.em mm.oh . emeoumeoue Aeeme oe.oe mm.mm He.em em.ee . emeoue-m Aeewv AeH.oev ANm.AmV Amo.emv Ame.ewv Ne.oe . om.em me.em wo.mw mmco exeemv me.--mmo he.oe mm.em hm.wm om.eoH emcoummo-e Acemv dale)». Nola i .mJIHIo. mm anew TC... 393 .Aemv mo Amze acne zeev humane sconceee mzz o m «Home ma 33 Table 4 JTINMR chemical shifts (PPM from TMS) of (27) Compound Aryl group C1,3 04,8, C5,7 Misc. 9,10a & C6 (27a) 4-CH30-06H4 4.220 2.700 2.288 CH3— 4.372 2.363 Ar-8.740(2H;d;j=9.5Hz) -7.424(2H;d;j=9.5hz) (27b) A-CH3-C6H4 4.438 2.811 2.311 CH3-2.842 2.469 Ar-8.698(2H;d;J=7.25Hz) -7.838(2H;d;J=7.25hz) (27c) 4—F-C6H4 4.459 2.888 2.333 Ar-7.680(2H;t;J=8.5hz) 2.546 -8.912-8.970(2H;m) (27d) C6H5 4.571 2.939 2.355 Ar-8.811(2H;d;J=8.3Hz) 2.590 —8.668(lH;t;J=8.3Hz) -8.022(2H;t;J=8.3Hz) (27e) 3-Br-C6H4 4.576b 3.001 2.365 Ar—8.879(1H;s) 4.528 2.643 -8.748(1H;d;J=7.8Hz) -8.646(1H;d;J=7.8Hz) -7.9l4(1H;t;J=8.0Hz) (27f) 4-CF3-C6H4 4.685 3.099 2.390 Ar-8.876(2H;d;J=8.3Hz) 2.726 -8.221(2H;d;J=8.3Hz) (27g) 3,5-(CF3)2-C6H3 4.745 3.202 2.430 Ar-9.l32(lH;s) 2.827 8.988(2H;s) (27h) 3,5-C12-4- 4.681 3.286 2.427 CH3-3.774(d;J=4.5Hz) N(CH3)2-C6H2 2.917 Ar-8.773(2H;s) (27i) 3,5-C12-4- 4.819 3.482 2.496 Ar-8.610(2H;s) CN-C6H2 3.061 2.459 8All of these peaks appeared as doublet with a coupling constant varying between 11.0 and 13.0 hz. bThe unsymmetrical aryl group causes these two hydrogens to become non- equivalent. 34 Nee (Misseoz m.am m.sm s.em e.Ne e.os~ (a- Ho-m.m Aesmv N m mmeo s.om o.sm ~.em m.mm N.emm - A movnm.m Amsmv m.om H.sm s.~m m.em c.mm~ emco-mmo-e Adamo m.om m.em m.Hm H.em H.5AN emeouamum Amsmv H.oM m.em m.se s.am m.ss~ memo AeeNV o.om o.am e.me ~.Hm N.me~ ameoum-e Acsmv s.e~-mmo m.m~ m.em m.me «.se m.oe~ emeoummoue Aesmo c.8mummo m.m~ m.em m.me m.ae 8.4mm emeo-0mmo(e Assay .emaz .mw same oH.s.m.eo maso mm. .mmcam Haws ecsowan .ANNV mo Amze comm ammo mousse accesses mzz 02 m mamas 35 36 The latter system, however, rearranged and consequently, no 13 C NMR 'could be obtained for it even at -110°C. To determine if this rearrange- ment was unique for the adamantyl system, the corresponding cyclopentyl cation (19) was generated and was found to be a complex mixture of products. No definitive spectra could be obtained for it. The 2-(p-anisy1)adamanty1 cation showed a peculiarity in its 13C NMR. All peaks showed a dependence on the concentration of acid used. This dependence is shown semiquatitatively in table 6. Tables 7 and 8 show the 1H NMR and 13 C NMR results respectively for the 7-aryl-7-norborny1 cations (28a-f). Olah47 has obtained chemical shifts for several of these systems and his results are included in table 8. The peak assignments for the exo and endo hydrogens were deduced by generating the 7-(p-fluorophenyl)norbornyl cation (28c) with deuterium placed on the exo face of the 2 and 3 positions. (exo, exo-2, 3—D2-2aé) The 7-(3,5-dichloro-4-N,N—dimethylammoniophenyl)-7-norbornyl cation (28g) could not be generated. All attempts to make it resulted in the formation of a system showing less than the expected symmetry of the classical cation. Furthermore, quenching the sample with sodium meth- oxide dissolved in methanol (five fold excess) at -80°C yielded the start- ing alcohol and not the expected methyl ether. To insure that the methyl ether was not cleaved to the alcohol under the work—up conditions, the o.mm q.wo ~.Ho .xmmm ecu mo mmosnmoun onu mo mmsmomn oumaaxoumem ma w.om m.om H.mN a.m~ H.5m 0.5m n.0m N.om h.mo o.mm n.0m N.oq m.n¢ oa.s.m.eo n.0m w.~m N.oq w.¢¢ m.Ho ¢.omm mm.eem «.mqm m.nm~ smash Hmeaamee No seem mucmam>fiovm o mmucmam>fisvm q mucmam>wsvo N uamHm>fisum H new: wave mo unseen mumafixouee< .GOHumuuaouaoo mace ou uoumwmu sues Amnmv.uo Amze Baum Snag mumwsm Hmoaaono mzz omH mo mocomcoeoo c canoe 37 .chfiuwmoe once was oxm ecu cu accuoue oaoaxnume ecu cwwmmm ou meme ma unsouum 02o . .o>Hum>wpom oueusevfivIMsN)OXe.oxm mnu mo muuowmm ecu Bouw monummcfi mums humane HmoHEoLo smacks .Nm m.m new w.m coosuon wcwmum> acoumaoo wswaesoo m saws mumfinsom mm venomous mxmma omens mo HH