A STUDY OF MOLECULAR STRUCTURE ‘ ANO INTERNAL ROTATION IN AMIDES .BY NUCLEAR MAGNETIC RESONANCE f . SPECTROSCOPY . * ‘ Thesis for the Degree Of Ph. D. MICHIGAN STATE UNIVERSITY LESTER REINHARDT ISBRANDT 1972 THE-9’9 This is to certify that the thesis entitled A STUDY OF MOLECULAR STRUCTURE AND INTERNAL ROTATION IN AMIDES BY NUCLEAR MAGNETIC RESONANCE SPECTROSCOPY presented by LESTER REINHARDT ISBRANDT has been accepted towards fulfillment of the requirements for PH.D. degree in CHEMISTRY Date /,” 12—" 7147 0-7639 BINDING BY IIIIMI & SDIIS' 800K BINDERY INC. LIBRARV BINDERS I LIBRARY Michigan Stat: University ABSTRACT A STUDY OF MOLECULAR STRUCTURE AND INTERNAL ROTATION IN AMIDES BY NUCLEAR MAGNETIC RESONANCE SPECTROSCOPY BY Lester Reinhardt Isbrandt High-resolution nuclear magnetic resonance (NMR) methods have been employed in a series of studies of molecular structure and internal mo- tions in amides. A Varian HA-IOO NMR spectrometer was used to obtain both proton and carbon-13 spectra. Activation parameters for internal rotation about the central C-N bond in six previously studied N,N-dimethy1amides, RCON(CH3)2 (R=H,D, -CH3,-CH2CH3,C1,C013), have been redetermined from proton NMR spectra by total lineshape analysis. The NMR method for carrying out the analy- sis has been improved by the digitization of the experimental spectra followed by computer curve fitting of the lineshapes to the theoretical lineshape equation containing all the NMR parameters as adjustable vari- ables. The agreement between the activation parameters determined in this study and the literature values obtained by total lineshape analy- sis is quite good and establishes the reliability of the method. Internal rotation about the central C-N bond in a series of four unsymmetrically N,N-disubstituted amides, CH3CON(CH3)R (R=ethyl, iso- propyl, Efbutyl, and cyclohexyl), have been obtained by the proton NMR total lineshape technique developed for the study of the N,N-dimethyl- amides. Steric effects appear to be of particular importance and larger Lester Reinhardt Isbrandt groups, whether substituted on carbon or nitrogen, consistently reduce the rotational barrier. Application of lanthanide shift reagents has been proven to be a reliable method of assigning the NMR resonances, proton or carbon-13, to substituents on the gi§_and Egggg_rotational isomers of disubstituted amides. Assignments have been made in the proton NMR spectra for a rep- resentative series of amides and in the carbon-l3 NMR spectrum of N,N—di- gyprOpylformamide. The carbon-l3 NMR chemical shifts of fifty monosubstituted and di- substituted amides have been measured. Assignments have been made by the use of model systems and the application of lanthanide shift reagents. The chemical shift difference between carbon atoms of groups £i§_and £5325 to carbonyl oxygen in disubstituted amides is greatest for thecx- carbon, 22, 5.1 ppm in formamides and ca, 3.5 ppm in alkylamides, and decreases for carbons farther away from the carbonyl group. A STUDY OF MOLECULAR STRUCTURE AND INTERNAL ROTATION IN AMIDES BY NUCLEAR MAGNETIC RESONANCE SPECTROSCOPY BY Lester Reinhardt Isbrandt A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1972 To Pamie ACKNOWLEDGEMENTS The author would like to express sincere appreciation to Professor Mex T. Rogers for his guidance, encouragement, and patience during the course of this study. The financial support of the Dow Chemical Company, the National Science Foundation, and the Chemistry Department during the various parts of this study are gratefully acknowledged. To Pamela, my wife, whose ever-present smile, encouragement, and self-sacrifice enabled me to complete this phase of my education, I offer my deepest appreciation. TABLE OF CONTENTS Page INTRODUCTION ..... ............. . .............. . .................... l HISTORICAL ........ ................. . .......... .... .......... . ..... 3 Determination of Barriers in Amides . .................... ..... 4 Assignments of Resonances in Amides ..... ................ ..... 8 Application of Carbon-13 NMR Spectroscopy to Amides .... ...... 9 Additivity of Carbon-13 Chemical Shifts ............ .......... 11 THEORETICAL . ................. .......... ........ . .................. 14 Introduction ................................................ 14 Spin-Spin and Spin-Lattice Relaxation ................. . ...... 15 NMR Lineshape Theory ..... .. ..... ........................ ... 16 The Lineshape Equation with Exchange ......................... 19 Activation Parameters ............... ........................ . 22 Chemical Shifts ........... . .............. .... ................ 23 Nuclear Spin-Spin Coupling ..... .................... ..... ..... 25 Proton Decoupling .... ...... . .... .......... . ................. 27 Nuclear Overhauser Enhancement .......... . .................... 28 Lanthanide Induced Chemical Shifts ........................... 32 EXPERIMENTAL ...................................................... 34 Preparation of Compounds ..................... ... ..... . ...... 34 Physical Constants of Compounds Studied ................. ..... 35 Spectrometer ................................................ 40 Time Averaging and Digitization of Spectra ................... 43 Sample Preparation .. ..... .... ..... ... ................... 43 Bulk Diamagnetic Susceptibility Corrections .......... . ....... 45 Determination of Energy Barriers of Hindered Internal Rotations ......................................... 46 RESULTS . ................................... . ..................... . 48 Hindered Internal Rotation in Tertiary Amides ................ 48 Symetrically Disubstituted Amides ......................... 48 N,N-Dimethy1acetamide ............ . ..................... 48 N,N-Dimethylcarbamylchloride ........................... 55 N,N-Dimethylformamide .................................. 55 N,N-Dimethylformamide-d1 ..... . ......................... 55 N,N-Dimethylpropionamide ...... ...... . .................. 62 N,N-Dimethyltrichloroacetamide ......................... 62 TABLE OF CONTENTS - Continued Page Unsymmetrically Disubstituted Amides ................. . ..... 62 N-Ethyl-N-methylacetamide ............ ......... . ....... . 62 N-nfButyl-N-methylacetamide ............................ 62 N-Cyclohexyl-N-methylacetamide ............. ............ 62 N-Isopropyl-N-methylacetamide ... ....................... 73 Resonance Assignments in the Proton NMR Spectra of Several Tertiary Amides . ....... .......... ....... ... ........ 73 l-Methyl-Z-pyrrolidinone .................... . .......... 73 N,N-Dimethylformamide ................ . ................. 78 N,N-Diethylformamide ................................... 78 N ,N-Diisopropylformemide ............................ ... 82 N,N- Dimethylacetamide .... ............................. 82 N,N- -Dimethy1carbamylchloride .............. ...... ....... 86 N ,N-Dimethyltrichloroacetamide .................. . ...... 86 N,N-Dimethyltrifluoroacetamide ..... . ................... 86 Ethyl-N,N-dimethylcarbamate ............ ......... . ...... 86 N-Isopropyl-N-methylformamide .......................... 91 Carbon-l3 Chemical Shifts of Aliphatic Amides ................ 91 Monosubstituted Amides ................................. 93 Symmetrically Disubstituted Amides ..................... 9S Unsymmetrically Disubstituted Amides ................... 100 DISCUSSION ......... ..... . ......... . ..... . ............ .... ......... 103 Hindered Internal Rotation ...... ............................. 103 Total Lineshape Analysis ..... .............................. 103 The Frequency Factor in the Arrhenius Equation .. .......... . 104 Rotational Barriers in N,N-Disubstituted Amides ........... . 105 Resonance Assignments in Tertiary Amides by Lanthanide-induced Shifts .................................. lll Carbon-13 Chemical Sh1fts of Am1des .......................... 115 SUMMARY .. .................................................... 119 LIST OF REFERENCES ....... . ....... . . ............... . ............. 120 ii LIST OF TABLES TABLE Page 1. Reported activation parameters for hindered rotation in N,N-dimethylformamide ..................... ...... . ......... 5 2. Reported activation parameters for hindered rotation in some other dimethylamides ........................ .......... .. 7 3. Substituent parameters correlating alkane carbon-13 chemical shifts in ppm ...... ....... ............. .... ....... .. 12 4. Shielding effects of some common substituents in aliphatic systems ...... ............ .... .................... . ........... l3 5. N-monosubstituted amides studied ........................... .. 36 6. N,N-Dimethylamides studied .................. . ................ 37 7. Other symmetrically N,N-disubstituted amides studied ......... 38 8. Unsymmetrically N,N-disubstituted amides studied ........... . 39 9. Volume magnetic susceptibilities and bulk magnetic susceptibility corrections for several compounds ... ....... ... 45 10. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylacetamide ..... .......... .................... 53 11. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylcarbamylchloride ............................ 56 12. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylformamide - Method I ................... ..... 58 13. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylformamide - Method II ............ ........... 6O 14. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylformamide-dl ................................ 63 iii LIST OF TABLES - Continued TABLE 15. 16. 17. 18. 19. 20. 21. 22. 23.- 24. 25. 26. Page Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N’N-dimethYIPropionamide ......OOOOOOOOOOOOOO......I... 65 Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N’N-dimethyltriChloroacetmide O I O O O O O O O O C O O O O O O I O O O O O O 67 Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N-ethYI—N-methYlacetamide coooooooocoo-0000000000000... 69 Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat NfErbuty1-N‘methYIacetamide oooooooooooooooooooooooooon 71 Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N'CYCthBXYl-N-methYIacetaMIde oooooooooooooooooooooooo 74 Experimental data and calculated activation parameters for the internal rotation about the central C—N bond in neat N-iSOpr0py1‘N-methy13C8tamide ooooooooooooooooooooooooo 76 Proton chemical shifts, observed and calculated, in 1- methyl—Z-pyrrolidinone with increasing amounts of Eu(f0d)3 0.00.0................0............OOOOOOOOOOOOOOOO 79 Proton chemical shifts, observed and calculated, in N,N-dimethylformamide with increasing amounts of EU(de)3 ......OOOOOOOOOIOOOOOO.....OOIIOIOOOOOCOO0.0.0.0... 8O Proton chemical shifts, observed and calculated, in N,N-diethylformamide with increasing amounts of EU(de)3 0............OOIOOOOOOOO.......OOIOOOOOOOOOOOOOIOO. 81 Proton chemical shifts, observed and calculated, in N,N-diisopropylformamide with increasing amounts of Eu(f0d)3 0.0.0.000...0.0.0.0000..............IOOIOOOOOOOOOOO 83 Proton chemical shifts, observed and calculated, in N,N—dimethylacetamide with increasing amounts of Euwuom vouuomom .N OHAMH 8 Although symmetrically N,N-disubstituted amides have been rather intensively investigated, relatively few measurements have been made on unsymmetrically substituted compounds. Rotational barriers in three formamides, HCON(CH3)R[R = mCH2-, 02CH- and mC(CH3)H-] were determined by Franconi (41) who found no significant difference among the three within the rather large errors involved. Gehring 35 a1. (38) reported activa- tion parameters for five compounds and showed that a large decrease in the barrier occurs when one of the substituents is the vinyl group. Gutowsky st 31. (29) made a study of N-methyl-N-benzylformamide, Weil _e_§ 31. (42) of N-methyl-N-picrylacetamide, and Mannschreck et al. (43,44) of N-mesityl substituted amides. Assignment of Resonances in Tertiary Amides Except in a few cases where there is chemical shift degeneracy, the ‘gig (to oxygen) and the £5333 substituents on nitrogen give well separ- ated NMR signals whenever rotation around the C-N bond is slow on the NMR time scale. The principal cause of the chemical shift between the gig and trans NMR signals is the anisotropy of the magnetic susceptibility of the amide grouping (46). The chemical shift is observed whether substi- tution on the nitrogen atom is symmetrical or unsymmetrical. Assignment of the observed resonances to protons at sites A,B has been of consider- able interest and is particularly important in the case of unsymmetrical substitution, when it serves to identify the isomers, III and IV (Figure 2). Also, the values of the proton chemical shifts for groups gig or ‘Egggg to oxygen yield information on the structure of the amide molecule, as well as the nature of the adducts and complexes that it forms with other molecules. 9 Several criteria have been used to identify the proton resonances for R1 and R2 as gig or 5533;. These include: (a) inequality of the gig and 55225 coupling constants (47) for the protons of R3 with those of R1 and R2, (b) differential solvent shifts in an aromatic solvent (48,49), (c) the nuclear Overhauser effect (50,51), and (d) contact shifts induced by complex formation with paramagnetic metal ions (124). Most of these methods are quite limited in scope of application and the assignments may not be unequivocal (51), particularly when R3,R1,R2 are larger groups. Application of Carbon-13 NMR Spectroscopy to Amides Very limited attention has been given to the carbon-13 NMR spectros- copy of amides compared to proton NMR since the low natural abundance (1.1%) of the carbon-13 isotope and the reduced magnetogyric ratio (7C/7N = 0.251) combine to greatly reduce the detectability of the carbon-13 NMR signal. The signal-to-noise ratio may be significantly enhanced with proton decoupling techniques, increased sample size, and utilization of time-averaging techniques. The instrumentation necessary for, and chemical applications of, carbon-l3 NMR spectroscopy have been recently discussed by Stothers (52) and by Randall (53). The first reported observation of the carbon-13 resonance in amides was made by Lauterbur (54), who observed the carbon-13 chemical shifts in N,N-dimethylformamide. McFarlane (55) reported the carbon-13 chemical shift difference between N-methyl carbons in N,N-dimethylformamide, N,N- dimethylacetamide, and N,N-dimethylcarbamylchloride in five different solvents of various dielectric strengths. The chemical shift differences were invariant to the choice of solvent. He pointed out that the lO principal contribution to the carbon-13 chemical shift differences be- tween the Ndmethyl carbons in these amides arises from an intramolecular electric field. This field will also affect the proton chemical shifts, although to a smaller extent, since it is the paramagnetic contribution to the shielding which is involved. Levy and Nelson (56) studied the spin-lattice relaxation times in N,N-dimethylformamide, N,N-difig-butylformamide, and N,N-di-Efbutylaceta- mide by carbon-13 Fourier transform NMR spectroscopy. They found sig- nificant carbon steric compression shifts for the four aliphatic carbons £5223 to the formyl hydrogen and eclipsing the carbonyl oxygen in N,N-di- ‘g-butylformamide. These steric shifts range from over 5 ppm for the a-carbon to ca: 0.1 ppm for the 8-carbon on the same chain. Carbon spin- lattice relaxation behavior in N,N-dijg-butylformamide indicated that the ends of both butyl chains have significantly increased motional freedom. Gansow, Killough, and Burke (31) studied hindered internal rotation in N,N-dimethyltrichloroacetamide by carbon-13 NMR spectroscopy. Their results are comparable to the best activation parameters by proton NMR. However, they pointed out several experimental shortcomings of the tech— nique. First, the reported spectra were the absolute limit of the carbon- 13 NMR sensitivity and several thousand scans were necessary. Second, as lines broaden or as measurements are performed at higher temperature, sensitivity drops dramatically for two reasons: (a) Boltzmann popula- tions readjust,and (b) as a result of the large low-temperature chemical shift difference, broadening and coalescence occur over a large spectral area. ll Additivity of Carbon-13 Chemical Shifts Grant and Paul (57) found that alkanes absorb over a range of approx- imately 45 ppm and also discovered that the chemical shifts could be des- cribed by the relation 1 5c =B+JgAjnllj , (1) where 8: is the ith carbon-13 shift, A an additive shift parameter for J the jth position, n is the number of atoms in the jth positions, and 1) B is a constant. For linear alkanes five parameters (Aj) correlate 30 shifts ranging over 37 ppm, with a standard deviation (0) of 0.2 ppm. For the branched alkanes additional parameters are required to account for the shieldings of highly substituted carbons and their immediate neighbors. With a total of 13 parameters, 53 measured chemical shifts are correlated with o = 0.3 ppm. Five of these parameters, labeled<2, B, 7, 8, and 5 factors, represent the effect of replacing a hydrogen atom in the indicated position with a methyl group while the remainder account for the effects of branching; for example, a factor for methyl carbons bonded to a tertiary carbon is denoted 1°(3°). These substituent param- eters are listed in Table 3. The constant B has the value -2.5 ppm. To illustrate the manner in which these parameters may be employed, consider C-2 of 3-methy1pentane for which 8c is given by [B +IZa + 20 + 7 + 2°(3°)] OBS = 29.5 ppm, while 5c = 29.3 ppm (57). Several families of substituted hydrocarbons have been examined by this empirical method and the general trends follow those found for hydrocarbons, with the additional feature that polar substituents exhibit 12 Table 3. Substituent parameters correlating alkane carbon-13 chemical shiftsa in ppmb. a 9.1 i 0.1 1°(3°) -1.1 i 0.2 s 9.4 2 0.1 1°(4°) -3.4 i 0.4 7 -2.5 x 0.1 2°(3°) -2.5 i 0.2 5 0.3 i 0.1 3°(2°) -3.7 i 0.2 e 0.1 i 0.1 4°(1°) -1.5 1 0.1 8Reference 57 bIn this and the following Table, a positive value indicates a shift to low field. Values from single observations: 2°(4°), -7.2; 3°(3°), -9.5; 4°(2°), -8.4. Symbols are defined in the text. rather larger inductive deshielding effects at the immediate neighbors (i.e., theIJ and 8 effects) while the 7 effects are comparable to those of methyl groups. To illustrate the effects produced by a selection of substituents, some representative results are given in Table 4 most of which were obtained from the data for l-substituted alkanes. Perhaps the most distinctive feature of these substituent parameters is the change for the effect of a 7-methyl carbon with respect to the<1 and 8 effects. The original proposal by Grant and Paul (57) that the 7 effects are caused by conformational interactions appears to be sound, since it is generally observed that carbon atoms in sterically congested environments tend to absorb at higher fields than those in otherwise comparable orientations. Table 4. substituents in aliphatic systems. 13 Shielding effects8 of some common bExcluding 2-a1cohols, RCHOHCH 3. ..9L. .Ji_. _JL_ ._§_ .jL. .Bsfgssggg 01 31.2 10.5 -4.6 0.1 0.5 58 Br 20.0 10.6 -3.1 0.1 0.5 58 I -6.0 11.3 -1.0 0.2 1.0 58 NH2 29.3 11.3 -4.6 0.6 0.6 53 . 1° 48.3 10.2 -S.8 0.3 0.1 59 on b 2° 40.8 7.7 -3.7 0.2 0.3 59 coon 20.9 2.5 -2.2 1.0 1.2 60 000' 24.4 4.1 -1.6 1.2 60 RCO 30 1 -2 1 61 aIn ppm. THEORETICAL Introduction Nuclear magnetic resonance is a spectroscopic phenomenon observed only for nuclei which possess a magnetic moment, it given by if = 7i“? = VIII/2n . (2) where 7 is the magnetogyric ratio of the nucleus, S is the angular momen- tum of the nucleus, h is Planck's constant, I is the nuclear spin vector, given bygI = I(I+l), and I is the spin quantum number. When the nucleus is placed in a uniform magnetic field, there are (21+1) energy levels available, each corresponding to a different component of the angular momentum with values I, I-l,...—I+l, -I. The absorption or emission of an appropriate quantum of energy, AB = hv = 7hHo/2fl = PHD/I 9 (3) will enable the nucleus to make a transition from one energy level to an adjacent level. In order to observe an energy transition the population of the energy levels must be different. For an ensemble of nuclei the relative populations are given by the Boltzmann factor 14 15 N l m“Ho ii— : exp[(21+l)mHO/IRT] :: “2'11?“ - m, ) , (4). where m is the nuclear magnetic quantum number with values I, I-1,... -I+l, -I, k is the Boltzmann constant, and T is the absolute temperature. After the nuclei have undergone a transition, some mechanism is required whereby nuclei in an upper spin state can relax to a lower state so that the absorption of energy may continue. Spin-Spin and Spin-Lattice Relaxation Spin-lattice relaxation is a process by which a nucleus in an upper spin state may give up energy to the lattice in the form of translational or rotational energy. Random molecular motions of magnetic nuclei result in fluctuating magnetic fields, which may have an oscillating component whose frequency will match the precessional frequency of the magnetic nuclei in the upper spin state. When the two are in phase, the nucleus will be able to lose its extra energy to the lattice and drop to a lower energy level. The rate that magnetic nuclei of spin I = 1/2 approach their equilib- rium distribution, no, is given by 3: T1 0 , (5) where T1 is the spin-lattice relaxation time, and n is the excess popula- tion of the lower energy state at time t. The magnitude of T1 depends on the magnetogyric ratio of the nucleus and on the nature of the molecular l6 motions. The reciprocal of T1 gives an approximate measure of the line width due to spin-lattice relaxation. Another type of relaxation is by spin-spin'interaction. The pre- cession of a magnetic nucleus about the fixed axis of a uniform magnetic field, fig, may be resolved into a static component parallel to E; and a rotating component parallel to BL. ‘When the rotating component is at the correct frequency, there is an exchange of spin energy between two neighboring nuclei. Spin-spin relaxation thus results in no net change of spin state. The spin-spin relaxation time is further decreased by the small variation in the local field at each nucleus due to the fields of neighboring nuclei. The small local magnetic fields will add to or subtract from the applied field, fi;, resulting in a larger range of fre- quencies at which nuclei of the same type will absorb energy. Rapid rotation and tumbling, as found in most liquids and gases, will average the local magnetic fields so that T1 and T2 become equivalent. NMR Lineshape Theory To describe the time dependent variation of the components of the total nuclear magnetic moment per unit volume, Bloch used a set of phenom- enological equations (62). Bloch considers a nucleus with III = (h/Zx)- TI(I:I) and magnetic moment fi== ngas a tiny gyroscope. The forces it experiences in an external constant magnetic field in the z direction, E:= [O,O,HO}, cause it to move in such a way that the rate of change is given by the torque 0.0. n Inez ll 2 2 713x13] . (6) 17 with components dux/dt = 7lusz - quy] = 71120), (6a) du/dt = ”“sz - utz] = -7qux (6b) duz/dt = 701x11), - uny] = 0 (6c) These equations indicate that the nuclear moment : precesses about the z-axis with an angular frequency) + 1/“12A + pB/TlGA + 7GB =-ipA7H1Mo (15a) PB . 1 . ? GA - [1((DB-(D) + /T2B + pA/T]GB = -1p B7H1Mo - (15b) These are easily solved if one defines GA = ~[2ni(vA-v) + l/T2A + pB/T] (16a) OB = -[2ni(vB-v) + l/TgB + pA/r] (16b) 0 = 7H1MO . <17) 21 where VA and VB are the chemical shifts in hertz relative to some standard and C may be taken as an arbitrary scaling factor; the total transverse magnetization, G = GAd-GB, is given by '1CTIZPAPB - 1(pAOtB + pBClAH G = . (18) pApB " TgaAo‘s The real part of Equation (18) gives the dispersion u mode lineshape function and the imaginary part gives the absorption v mode lineshape function over the entire sweep range v as a function of the parameters VA, VB, T2A’ T23, pA, pB, and 1. These functions maybe written as (11) 'C{QP " [1+ 1(pB/T2A + pA/T‘QB) 1R] u = P2 + R2 (19a) -C{P[l+r(pB/T2A + pA/TQB)] + QR} v = P2 +-R2 . (19b) The quantities P, Q, and R are defined for convenience as P = 'r[l/T2AT23 -(2n)2Av2 + (2fl)2(5v/2)2] + pB/T2B + pA/TgA (20a) Q = 2mm - €35! (pA - 93)] (20b) R = 2nAw[14-T(l/T2AeFl/T23)] + n18v(l/TgB-l/T2A) (20c) where Aw = v-vo, Va is the average resonance frequency of VA and VB’ and CV = vA- VB. 22 Activation Parameters The activation energy for the exchange process in N,N-disubstituted amides can be evaluated from the Arrhenius rate equation which is written as (65,66) k = A exp(-Ea/RT) , (21) where k is the rate constant which is equal the inverse of twice the mean lifetime (1), R is the molar gas constant, A is the frequency factor, and E8 is the activation energy for internal rotation or the energy barrier of the system. The quantities A and E A least-squares analysis from the experimental data for each amide. can be evaluated by a linear If one assumes that this exchange process obeys the absolute rate equation (66), the entropy and enthalpy of activation, AS* and AH}, for the internal rotation process can be written as kBT * i k = K -fi-exp(-AH /RT) exp(AS /R) , (22) where r is the transmission coefficient, which is unity when every acti- vated complex breaks up to give products, kB is the Boltzmann constant, and h is Planck's constant. The free energy of activation at temperature T°K can be found from the relationship AS = AH - TAS . (23) 23 Chemical Shifts Chemical shifts, for molecules without unpaired electrons, have their origin in the magnetic screening of the nucleus which arises from the orbital electronic currents induced by an external magnetic field (67). These currents also give rise to diamagnetic polarization. When the external magnetic field is Ho the total magnetic moment of the induced current is XMHo where XM is the molecular diamagnetic susceptibility. The secondary magnetic field due to the induced currents at any given nucleus is -01Ho where 01 is the magnetic screening constant. The total field experienced by a given nucleus, which determines its NMR frequency, is given by (68) N1 = Ho(l-oi) . (24) The chemical shift may then be defined as Hi - I'Iref 8 (ppm) = -—fi—'—-— X 106 , (25) ref where H1 is the resonant field of resonance being measured at a fixed frequency and Href is the resonant field for a given reference. The chem, ical shift may also be defined in terms of frequency as vi - vref 6(ppm) = --;—-——— x 106 , (26> ref where v, is the resonant frequency of a signal being measured at a fixed field and vref is the resonant frequency for a given reference. 24 From Equation (24) it is seen that the differences in the shieldings of various nuclei are reflected by the differences in the screening con- stants. The theory of screening constants first proposed by Saika and Slichter (60) gives an atomic breakdown of diamagnetic currents. The total screening constant for any atom A may be broken down into the fol- lowing four terms: 0 =0 ‘1 a A AA + OIAA +21 UA13 + UA,r1ng . (27) d 1304A) The first term, 02A, field at nucleus A due to diamagnetic currents on atom A itself. For an is the contribution to the secondary magnetic isolated spherical atom, it is the only contribution and is given explicitly by the Lamb formula, AA e2 -1 Cd = 3mc2 E: ri ’ (28) where e is the charge of the electron, m is the electron mass, c is the velocity of light, and r is the distance of the ith electron from the i nucleus, the sum being over all the electrons. The second term, ogA, is the contribution due to the paramagnetic currents on atom A, which gives the paramagnetic susceptibility. This term was first calculated by Saika and Slichter who showed that for fluorine atoms it was much more sensitive to chemical structure than 03A . For most other nuclei, variations in this term will give the main contribution to the chemical shift. The principal exception is hydrogen, where the absence of low-lying atomic p-orbitals will make the paramagnetic term negligibly small. 25 The third term, GAB, is the contribution to the screening of atom B ) A by the atmmic circulations on atom B. When the magnetic effects of these neighboring currents are treated in a dipole approximation, this term involves only the local anisotropy of the local susceptibility on atom B. If atom A is on, or near, the axis of high diamagnetism of B, there will be an increase in the average screening. The effect falls off as the inverse cube of the AB distance. A,ring . Finally, 0 is the contribution of the screening due to ring currents which cannot be localized on any atom. The magnitude of this term is usually small, except it does play an important part in deter- mining the proton spectra of aromatic compounds. For carbon-l3 chemical shifts the last two terms are relatively unimportant, and calculations of the diamagnetic term, 0AA d 9 is fairly constant for different types of carbon atoms. The paramagnetic show that it term, 02A , is therefore the dominant term in the expression for the calculation of the shielding for the carbon-13 nucleus (70). Considerable attention has been paid to the evaluation of this term using quantum mechanical treatments (71,72). Nuclear Spin-Spin Coupling In 1951 Gutowsky, McCall, and Slichter observed that high resolution Spectra frequently exhibited hyperfine structure which, in contrast to the linear dependence of chemical shifts, was independent of the applied mag- n-‘E!tic field (73,74). Similar effects had been observed by Hahn and Maxwell it! the modulation of the spin-echo envelope in the pulse experiments (75). 1‘: was concluded that the observed effects were caused by an indirect 26 interaction of the nuclear moments pi, which is transmitted from nucleus to nucleus by the paired electrons comprising the valence bonds. The magnitude of this interaction is called the spin-spin coupling constant,.L The interaction can be theoretically divided into three parts (76). In the first of these, one nuclear magnetic moment induces orbital elec- tronic currents which consequently induce magnetic fields at the site of the second nucleus. In the second, the electron spin interacts with the magnetic moment of one nucleus producing an electron spin polarization which is then transferred to the adjoining nuclei. The third part, and the most significant contribution to the overall spin-spin coupling, is the Fermi contact interaction between nuclear spins and the spins of electrons in s-orbitals, which produces an electron spin polarization pro- portional to the density of the electrons at the nucleus. In molecular orbital theory the spin-spin coupling constant, JAB’ may be written in terms of the wave functions of the ground and excited states, W1 and W (77). Summing over occupied and unoccupied levels for J i and j, respectively, OCC. UUOCC. (I! ‘1’.) (‘II ‘1’) - 1.1 A iAj B JABa Z Z (G -<-: ) ’ (29) 1 j j 1 -e1) is the energy difference between the occupied ground state where (c J 6 and unoccupied excited states and is always positive. From the 6,, above expression the sign of the coupling constant depends on whether the i unlecular orbital is symmetric or asymmetric. The symmetric molecular orbitals have the same sign at atom A and atom B, and the asymetric molec- ular orbitals have different signs. For transitions from symmetric to symmetric or from asymmetric to asymmetric molecular orbitals, the 27 contribution to JAB is negative. Transitions from symmetric to asymmetric molecular orbitals, or vice versa, gives a positive contribution to the coupling constant. Proton Decoupling Spin decoupling, for the case where the coupled nuclei are of dif- ferent species, involves the application of a second radiofrequency field oscillating at the resonance frequency of the nucleus to be decoupled. This causes the nucleus to change spin states so rapidly that its spin vector no longer couples with the spin vectors of neighboring nuclei and multiplets due to this coupling collapse. In a simple AX molecule, the elimination of proton-induced splittings can only be expected to collapse a doublet into a singlet, giving an en- hancement of two-fold. Collapse of splittings in more complex molecules such as benzene can result in a signal-enhancement figure approaching an order of magnitude (78). The carbon-l3 magnetic resonance spectrum of benzene (79) exhibits considerable proton splitting beyond the large doub- let due to the directly bonded proton. These splittings are due to the long-range couplings between a carbon-13 and the remaining hydrogens in the molecule. The elimination of such complex splitting patterns leads to a dramatic improvement in the signal-to-noise ratio. As most organic molecules contain an even greater number of protons, benzene is by no means an exception to the rule, and multiplet collapse can be considered to give rise to about a lO-fold enhancement. It should be Pointed out that such techniques eliminate much useful information in the carbon-l3 spectrum, such as the coupling constants 28 between carbon and hydrogen. However, it may also be argued that the chemical shift information obtained under proton decoupling conditions is more readily acquired and interpreted. Elimination of proton split- tings leads to spectral singlets which are free from all second-order changes in line positions, and therefore shift data are obtained directly from.the position of the spectral singlet. Also, the number of directly bonded, geminal, and vicinal protons in most organic molecules will produce splitting patterns of the carbon-13 resonance signal much more complicated than is found in the corresponding proton resonance bonds. Thus, it is concluded that the loss of spectral information may in many instances be more than offset by the benefits of spectral simplicity, which allows the chemical shift information to be obtained more readily . and, in general, on a larger number of chemical systems. In addition to the enhancement due to multiplet collapse, proton decoupling techniques give rise to a nuclear Overhauser enhancement. Nuclear Overhauser Enhancement When the protons in a sample are decoupled by a strong irradiating field, the effect on the carbon-l3 NMR spectrum is an enhancement of the carbon-l3 signal due to a phenomenon known as the nuclear Overhauser effect (NOE). The NOE results from dynamic polarization of the carbon-13 nuclei when the proton resonances are saturated. This effect has been well studied in carbon-l3 (80,81,82) and in nitrogen-15 (83) NMR. The energy level diagram of a general two-spin system in a magnetic field is given in Figure 3 where the energy levels in the absence of a radiofrequency field have a Boltzmann distribution. When one nucleus is 29 Ba Figure 3. Energy-level diagram for an AX two-spin system in presence of a magnetic field. 30 irradiated by a radiofrequency field the other nucleus is enhanced by the nuclear Overhauser enhancement, 6, of 9 = l + nA-{x} , . (30) where = (yx/yA) I bGiven as: °C/mm. cStudied without further purification. 38 Table 7. Other symmetrically N,N-disubstituted amides studied. Compound N,N-Diethylformamide N,N-Diethylacetamide N,N-Diethylpropionamide N,N-Diethyl-nybutyramide N,N-Diethylchloroacetamide N,N-Diethylacrylamidec N,N-Diisopropylformamide N,N-Diisopropylacetamide N,N-Diisopropylpropionamide N,N-Di-gfpropylformamide N,N-Di-nypropylacetamide a Source 45. 60. 53. 61. 72 69. 56. 62. 60. 68. Boiling Point 0/48.0/2 5-61.5/2 0-54.o/1.5 0-63.0/1 .0-73.0/2 o-71.0/7 0/1 0-62.5/0.5 5-61.5/2.5 0-68.3/5 0060:» I bGiven as: °C/mm. cStudied without further purification. Matheson, Coleman and Bell, East Rutherford, New York. Eastman Organic Chemicals, Rochester, New York. K&K Laboratories, Inc. , Jamaica, New York. Prepared by L. A. LaPlanche in this laboratory (107). 39 Table 8. Unsymmetrically N,N-disubstituted amides studied. Compound N-Ethyl-N-methylformamide N-Ethyl-N—methylacetamide N-Ethyl-N-methylpivalamide N-Isopropyl-N-methylformamide N-Isopropyl-N-methylacetamide anfButyl-N-methylformamide Nfingutyl-N-methylacetamide Nfngutyl-N-methylisobutyramide anfButyl-N-methylpivalamide Nfithutyl-N-methylformamide NngButyl-N-methylacetamide N—Cyclohexyl-N-methylacetamide l-Methyl-Z-pyrrolidinone Source >* >' >- >' I> a Boiling Pointb 82.0/44 45.0/33 62.0-63.0/5 58.5/7 60.0/17 55.0-56.0/1 65.0/3 72.5-73.0/2.5 75.0-76.0/2 64.5/5 56.5/5 130/13 202/760 aA - Prepared by L. A. LaPlanche in this laboratory (107). B - Eastman Organic Chemicals, Rochester, New York, C - Aldrich Chemicals, Milwaukee, Wisconsin. bGiven as: °C/mm. 4O Spectrometer In this investigation, spectra were obtained on a Varian HA-lOO high- resolution nuclear magnetic resonance spectrometer. The main magnetic field of 23,490 gauss was generated by a V-4014 twelve-inch high-impedance electromagnet equipped with a V-2100B regulated magnet power supply. Stability of this magnetic field was maintained to better than one part in 108 by a V-3506 flux stabilizer. The system is capable of observing either proton resonances at 100 MHz or carbon-l3 resonances at 25.1 MHz. In each case, the crystal- controlled radiofrequency is produced by a separate V-4311 fixed frequency rf unit. The radiofrequency is transmitted to a V-4333 variable tempera- ture probe where the resonance of the sample is detected by the crossed- coil method. The carbon-l3 probe was double tuned so that the carbon resonances could be examined while irradiating the proton resonances to remove the coupling between the two nuclei. The pseudo-random noise at 100 MHz was generated by a V-3512 noise decoupler. A V-4354A internal reference NMR stabilization unit modulated the radiofrequency with two audio frequencies; one frequency, called the analytical channel, was used to detect the resonances in the sample and the other frequency, called the control channel, was used to lock the magnetic field to a reference frequency. The reference was one of the sample resonance frequencies or that of an added internal standard. Frequencies were counted with a Hewlett-Packard 5245L frequency counter. In order to observe a carbon-l3 spectrum it was necessary to work at the highest gain of the spectrometer. Working at this ultimate limit of the spectrometer produced an undesireable rolling baseline, as shown in 41 Figure 4, as the result of a large modulation index. The modulation index (110) is a dimensionless quantity, the ratio between the frequency deviation, in hertz, and the modulation frequency, also in hertz. When the audio frequency is swept, the modulation index is large and the am- plitude of the modulation signal varies greatly. The phase-sensitive detector used to detect the NMR signal is sensitive to changes in the modulation signal, which is normally small under the usual operating con- ditions of the spectrometer. However, under the conditions employed for the carbon-l3 NMR experiment these changes are very large and are re- flected in the spectrum. In order to conduct the variable temperature studies it was neces- sary to stabilize the temperature of the probe and measure this tempera- ture accurately. Temperature stability was maintained in the range of ~60 to 200°C by a V-4343 variable temperature accessory connected to a heater sensor in the probe. Temperature calibration for the proton work was accomplished by using the chemical thermometers methanol and ethylene glycol. In the high temperature range, 30°C to 200°C, the temperature was found from the expression (109) T°K = -(8EG) 1.017 + 464.9 , (38) where SE is the chemical shift difference between the two chemical shifts G of the protons in ethylene glycol at 100 MHz. In the low temperature range, -100° to 30°C, the calibration equation is (109) T°K = -(8M) 1.076 + 464.7 , (39) 42 .ooaamcowoouaahnuofiolz.z mo aouuooam mzz manoooumo .0 madman Eco . may“ 0. n. on am On an on no . c N: _ _ _ _ In a _ _ . _ 43 where 5M is the difference between the two chemical shifts of the protons in methanol at 100 MHz. The temperature calibration for the carbon-l3 spectra was accomplished by using a copper-constantan thermocouple be- cause no carbon-13 chemical thermometer has been found. Time Averaging and Digitization of Spectra The Varian HA-lOO was interfaced with a Varian C-1024 computer of averaged transients (CAT). When a spectrum of interest is swept repeat- edly, signal information is added into memory channels of the C-1024 in direct proportion to the number of sweeps, while random noise is accumu- lated in proportion to the square root of the number of sweeps. The sen- sitivity of the NMR experiment, especially for carbon-13 spectra, is increased by accumulating repeated sweeps of the spectrum. Frequency calibration is accomplished by reading the information stored in memory out on to a recorder precalibrated to the spectral sweep width. The C-1024 was also used to digitize a spectrum of interest for which purpose it was connected to a Varian C-1001 binary-to-octal coupler attached to a IBM 526 key punch. The punched cards containing the 1024 points in an octal format were read into a CDC 6500 computer for further processing. The C-1024 operated exclusively in this mode for the total lineshape analysis portion of this investigation. Sample Preparation All proton and carbon-13 spectra were obtained by placing the samples in five millimeter, precision-ground, NMR tubes. 44 The tertiary amides for the hindered internal rotation studies were degassed and sealed by the freeze-pump-thaw method to remove any trace amounts of paramagnetic oxygen which broadens the NMR lines. The addi- tion of five percent by volume of either tetramethylsilane (TMS) or hexa- methyldisiloxane (HMDS) was required in each sample to provide a refer- ence and lock signal. The substituted formamides, which have very high coalescence temperatures, were degassed and sealed in three millimeter tubes and placed in five millimeter NMR tubes containing tert-butyl benzene (TBB). The methyl resonance of TBB provided the reference and lock signal. Solutions of the tertiary amides used in the resonance assignment study were initially 0.2 M in carbon tetrachloride and the shift reagent [Eu(fod)3 or Pr(fod)3] was then added in increments up to a mole ratio of 0.4 (shift reagent to amide). Five percent by volume of TMS was added to each sample as an internal reference. The proton NMR spectrum was examined prior to and after each addition of the shift reagent. The series of spectra for each amide were acquired at a temperature well be- low the coalescence temperature._ The carbon-13 NMR spectra were obtained by placing the sample in a five millimeter NMR tube with a capillary containing 57% carbon-l3 en- riched methanol. The capillary was centered by means of a Teflon plug. The enriched methanol was used as a reference and lock signal. When necessary the spectra were run at temperatures low enough to slow the internal rotation and permit observation of both isomers. 45 Bulk Diamagnetic Susceptibility Corrections Where an external reference is used, a correction involving the difference between the bulk diamagnetic susceptibilities of the reference compound and the sample must be applied. This is necessitated by the fact that, in cylindrically shaped containers, the actual fields experi- enced by individual nuclei will depend on the magnetic polarization near the surface. ,The correction for a cylindrical sample is 2n SCORR ..— E’0133 + '3— (Xv,ref ' Xx) ’ (40) where 8 is the chemical shift in ppm and Xv is the volume diamagnetic susceptibility (68). The volume diamagnetic susceptibilities of several compounds were measured by the Gouy method (111). The measurements were made on a mag- netic susceptibility system manufactured by the Alpha Scientific Company. The measured values are given in Table 9. The chemical shift corrections Table 9. Volume magnetic susceptibilities and bulk magnetic susceptibility corrections for several compounds 2x Compound - X T (Xv,CH30H - Xv) Methanol 0,525 ppm ._ N-Methylformamide 0.564 ppm 0.081 ppm N-Methylpropionamide 0.619 ppm 0.197 ppm N,N-Dimethylformamide 0, 58 1 ppm 0. 117 ppm N,N-Diethylformamide 0,579 ppm 0,113 ppm N,N-Diethylacetamide 0.615 ppm 0.188 ppm 46 \ for several amides (with methanol aA\the reference) are also given in this table. The correction factor for formamides is approximately 0.1 ppm and for other amides it is about 0.2 ppm. Since these corrections are within the experimental error of measurement of the chemical shifts in the carbon-l3 spectrum, no bulk diamagnetic susceptibility corrections ‘were made for the carbon-13 chemical shifts. Determination of Energy Barriers for Hindered Internal Rotations The rate of internal rotation about the C-N bond in several amides was obtained by an iterative computer fitting of the theoretical spec- trum, calculated from the total lineshape equations as derived by Rogers and Woodbrey (11), to the experimental spectrum. The curve fitting program was developed by Dye and Nicely (112). The program minimizes the functional n q) =ZW1F12 , (41) i=1 in which n is the number of data points, F1 is the residual defined in such a way that it would approach zero for all i as the parameters approach their "best" values if the data were completely free from errors, and W1 is a weight which was set to unity for all points; if desired Wi can be calculated by the program. The general program will fit an arbi- trary function of not more than 20 parameters (dependent plus independent) to a data set containing a maximum of 99 points. The lineshape equation is a function of the spin-spin relaxation times at sites A and B (T2A’ T23), the chemical shift difference (8v) in 47 the absence of exchange, the relative population at each site (pA, pB), and the mean lifetime (1) at each site. These parameters were determined above and below the coalescence temperature for the Ndmethyl protons in each amide by the following techniques. The proton NMR spectrum of each amide was examined at a temperature well below the coalescence temperature to establish an estimated value of each parameter (except 1). In the temperature region below coalescence all of the parameters were deter- mined by simmltaneous iteration to obtain the best fit to the experimen- tal lineshape. The final value of each parameter calculated was examined for experimental feasibility. For temperatures above coalescence, the experimental lineshape contains insufficient information to allow the simultaneous determination of all the parameters. In this case, all of the parameters except the mean lifetime (1) at each site were held con- stant at a value which was the average of those calculated from the experi- mental curves below the coalescence temperature. RESULTS Hindered Internal Rotation in Tertiary Amides rev. The barrier to internal rotation about the central C-N bond for a i series of tertiary amides, symmetrically and unsymmetrically substituted, % 7 was studied by total lineshape analysis. The proton NMR spectrum of the I N-methyl protons was obtained at several temperatures above and below ’3 the coalescence temperature. The experimental curve at each temperature was then fitted to the theoretical lineshape equation in order to extract the parameters of interest. The proton NMR spectrum for each amide can be found elsewhere (107,108,109). Symmetrically Disubstituted Amides N,N-Dimethylacetamide --- The barrier to internal rotation for neat N,N-dimethylacetamide was studied at various temperatures by fitting the experimental curve of the N-methyl resonances to the theoretical line- shape equation. The parameters in the theoretical equation were then adjusted by an iteration procedure to obtain the fit. Four of these plots are shown in Figures 5-8. In these figures, X's designate experimental points, 0's designate calculated points, and ='s designate an experimental point and a calculated point which are in the same area element delta x by delta y. The experimental data obtained are given in Table 10. The loga- rithm of the rate constant is plotted against 103/T°K in Figure 9. 48 49 .Mem.m~m R 9 up oofismuoomamnuoafivuz.z use: aw mdououo Hhsumsuz new no mamawfim A00 moumaooamo mam Axv oo>uomno .m whomwm o 0800 0 x xxx 0 ouuu..x XXX 0 nun X .8 8800000000 3 x0 «an xxxxxx xxxxx 0 o n x O: x x0 «x o u x x0 4 x 4 x0 0 n X X o a o X I X u 0 n X n o _ x x A o x o x o u x o x x o o x x 0 o x o x 0 x x x o o 0 u o u Ox x 50 .MoH.o¢m I H. um ngumumfihfiumafifi Z. I 2 anon 5 3398 130812 on”. no 36:36 80 33333 one 03 350.30 .0 ousmwm OX 0 X n u OX 00 X0 X0 51 erode n e um SESSSESEEAE use: :a moououm Hanuoalz man no mamcwfim A00 omumaooamo 0am Axv om>uomno .n ouowfim OX X0 .0 OX X 52 rib .Mom.wcm 6H maououo HanuoEIz onu mo N H on unassumooahguoaaeuz.z umoa mamawwm A00 ooumasoamu can Axv oo>uomno .w ouswfim M 00 u uuquuuuquuuuuunuu O 53 Table 10. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylacetamide.i r°1< 103/ Tog 1 1n( 1/2.) 517* .299... _______._ .112. FA 325.7 3.071 .1181 1.443 17.7 .498 331.8 3.014 .0608 2.106 17.8 .503 335.1 2.984 .0419 2.480 17.6 .505 335.1 2.984 .0393 2.544 17.8 .496 337.3 2.965 .0331 2.716 17.2 .489 340.1 2.940 .0244 3.021 17.7 .503 343.0 2.916 .0213 3.157 17.5 .501 346.2 2.888 .0162 3.431 17.8 .500 348.5 2.870 .0138 3.586 350.8 2.851 0.0113 3.788 354.4 2.822 0.0084 4.082 358.1 2.792 0.0067 4.317 362.3 2.760 0.0051 4.594 E8 = 19.7 i 0.5 kcal/mole logloA = 13.9 i 0.4 AG*zgg= 18.1-i 0.5 kcal/mole 0H5 = 19.0 i 0.5 kcal/mole 43* = 2.9 i 1.4 eu Coalescence temperature = 348°K -* Results from total lineshape analysis. lav/2r) 4&6C" 4&20 3m30 3040 3K“) 126C) ELZO |.80 1440 54 .. O 1 n I J l 1 1 J 279 2.93 2.87 2.9: 2.95 299 3.03 3.07 Io3/T°K Figure 9. Plot of ln(l/27) against 103/T°K for neat N,N—dimethylacetamide. 7'] —-.—‘—v- “‘4'; ‘1 v—- v—- — L 55 NLN-Dimethylcarbamylchloride --- Hindered internal rotation in neat N,N-dimethylcarbamylchloride was studied by total lineshape analysis. The experimental data and calculated activation parameters are given in Table 11. The logarithm of the rate constant is plotted against 103/T°K in Figure 10. . ELN-Dimethylformamide --- The proton resonances of the N-methyl pro- tons in N,N-dimethylformamide are split into doublets due to coupling with the formyl proton. The trans coupling constant is 0.8 Hz and the gig coupling constant is 0.5 Hz. The barrier to internal rotation for this amide was studied by two methods. Method I. The coupling to the formyl proton was neglected and the experimental curves were analyzed as a two-site case. The experimental activation parameters calculated by this method are given in Table 12. The logarithm of the rate constant is plotted against 103/T°K in Figure 11. Method 11. The coupling to the formyl proton was taken into consid- eration by assuming the pair of doublets to be a consequence of the super- position of two doublets (109) whose centers are separated by (Jci + S Jtrans)/2' The experimental spectrum was then fitted to the theoretical lineshape equation which had been modified to include the coupling con- stants. The experimental data and the activation parameters calculated are given in Table 13. The logarithm of the rate constant is plotted against 103/T°K in Figure 12. NLN-Dimethylformamide-dl --- The effect of coupling by the formyl proton was removed by replacing this proton with a deuteron. The hindered internal rotation was studied by total lineshape analysis and the experi- mental data and the calculated activation parameters are tabulated in 56 Table 11. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylcarbamylchloride. T°K 103/T°K 1* 1n(1/21) 8v* pA* sec _§§_ 303.8 3.292 0.1623 1.125 10.8 0.498 308.7 3.239 0.0921 1.692 10.7 0.494 311.4 3.211 0.0746 1.903 10.8 0.502 316.0 3.165 0.0472 2.360 11.0 0.503 316.4 3.161 0.0481 2.341 10.7 0.492 319.3 3.132 0.0312 2.733 11.1 0.497 321.3 3.112 0.0240 3.036 11.5 0.506 323.5 3.091 0.0214 3.150 325.6 3.071 0.0198 3.230 328.8 3.041 0.0141 3.568 330.7 3.024 0.0130 3.647 E8 = 16.7 i 0.6 kcal/mole logloA = 13.9 i 0.4 063293 = 17.1 i 0.6 kcal/mole AH3 = 18.0 i 0.6 Real/mole * + AS = 3.0 - 1.8 an Coalescence temperature = 324°K * Results from total lineshape analysis. luv/21') 57 4.20 " 3.40 3.00 2 .60 2.20 I .80 L40 1.00 - 11111111 3.02 3.06 am 3.:4 3.Ia 3.22 3.26 103/T’K Figure 10. Plot of 1n(l/21) against 103/T°K for neat N,N-dimethylcarbamylchloride. 58 Table 12. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylformamide - Method I. 1--'--- -- -"‘—~ '- -~'y 4‘. K3 l T°K 103/T°K 7* 1n(1/21) 0v* pA* sec Hz 383.2 2.609 .0499 2.304 15.0 0.500 386.3 2.589 .0381 2.575 14.9 0.498 387.3 2.582 .0357 2.639 15.1 0.505 389.3 2.568 .0312 2.773 14.9 0.503 391.7 2.553 .0276 2.897 15.3 0.509 392.7 2.546 .0257 2.968 15.2 0.504 394.4 2.535 .0240 3.036 15.1 0.496 394.8 2.533 .0228 3.089 15.2 0.506 395.4 2.529 .0218 3.134 15.4 0.510 395.7 2.527 .0202 3.207 14.9 0.501 397.2 2.519 .0196 3.241 398.2 2.511 .0174 3.359 399.4 2.504' .0171 3.378 E8 = 19.8 i 0.5 kcal/mole 10g1oA = 12.3 i 0.4 AG:t = 20.4 i 0.5 kcal/mole Afli = 19.0 i 0.5 kcal/mole 43* = -4.7 t 1.5 eu Coalescence temperature = 397°K * Results from total lineshape analysis. 59 . c u 3.70 _ H’ \cu, 350 - 3.30 r- : 3.:0 — s“ 5. 2.90 - 2.70 - 2.50 - 2.30 — l 1 I 1 l 1 2.50 2.52 2.54 2.55 2.58 2.60 ' I03/7°K Figure 11. Plot of 1n(1/21) against 103/T°K for neat N,N-dimethylformamide - Method I. Table 13. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylformamide - Method II. 60 T°K 10?/T°K 1* 1n(1/21) 5v* pA* 229.. Hz 386.3 2.589 .0469 2.367 14.9 0.502 387.3 2.582 .0433 2.450 14.9 0.505 389.3 2.568 .0368 2.610 14.8 0.498 391.7 2.553 .0285 2.866 15.2 0.509 392.7 2.546 .0275 2.901 15.1 0.497 394.4 2.535 .0240 3.037 15.2 0.507 394.8 2.533 .0242 3.030 15.4 0.510 395.4 2.529 .0222 3.115 14.8 0.498 395.7 2.527 .0219 3.127 397.2 2.518 .0197 3.232 398.2 2.511 .0174 3.360 399.4 2.504 .0170 3.380 Ea - 24.3 .5 kcal/mole logloA = 14.8 i 0.4 £03293 3 21.5 1 AB = 6.5 i 2. i 0.5 kcal/mole 0 eu Coalescence temperature = 397°K ¢ AH = * Results from total lineshape analysis. 23.5 i 0.5 kcal/mole in (V2?) 3.70 3. 50 3.30 3.I0 2.90 2.70 2.50 2.30 61, P 1 1 1 1 1 1 2. 50 2.52 2.54 2. 56 2.58 2.60 Io3/T°K Figure 12. Plot of ln(1/2T) against 103/T°K for neat N,N-dimethylformamide - Method II. 62 Table 14. The logarithm of the rate constant is plotted against 109/T°K in Figure 13. ELM-Dimethylpropionamide --- Hindered internal rotation in N,N- dimethylpropionamide was studied in the neat liquid by total lineshape analysis and the experimental data and the calculated activation paramp eters are tabulated in Table 15. The logarithm of the rate constant is plotted against 103/T°K in Figure 14. M,N-Dimethyltrichloroacetamide --- Hindered internal rotation in N,N-dimethyltrichloroacetamide in the neat liquid was studied by total lineshape analysis and the experimental data and the calculated activa- tion parameters are tabulated in Table 16. The logarithm of the rate constant is plotted 103/T°K in Figure 15. Unsymmetrically Disubstituted Amides N-Ethyl-N-methylacetamide --- Hindered internal rotation in N-ethyl- N-methylacetamide in the neat liquid was studied by total lineshape analy- sis of the N-methyl proton resonances. The experimental data and the calculated activation parameters are tabulated in the Table 17. The log- arithm of the rate constant is plotted against 103/T°K in Figure 16. N-EfButyl-N-methylacetamide --- Hindered internal rotation in N-nf butyl-N-methylacetamide in the neat liquid was studied by total lineshape analysis of the N-methyl proton resonances. The experimental data and the calculated activation parameters are tabulated in Table 18. The logarithm of the rate constant is plotted against 109/T°K in Figure 17. N-Cyclohexyl-N9methy1acetamide --- Hindered internal rotation in N— cyclohexyl-N-methylacetamide in the neat liquid was studied by total 63 Table 14. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylformamide-dl. T°K 103/1°K 1* 1n(1/21) 59* pA* _§2£._ __.___... Hz 370.3 2.700 0.279 0.583 15.0 0.506 373.9 2.675 0.217 0.835 15.2 0.502 379.0 2.639 0.118 1.446 15.0 0.499 383.5 2.607 0.0805 1.827 15.0 0.503 388.6 2.573 0.0651 2.038 14.9 0.496 397.8 2.514 0.0243 3.025 15.0 0.497 403.9 2.476 0.0169 3.390 15.1 0 501 407.5 2.454 0.0119 3.737 412.5 2.424 0.0082 4.105 417.1 2.397' 0.0056 4.491 422.2 2.369 0.0043 3.749 Ea = 25.3 i 0.3 Real/mole logloA = 15.2 1 0.2 AG*298 = 22.1 i 0.3 kcal/mole AH3 = 24.6 i 0.3 kcal/mole 28* = 8.5 i 1.2 an Coalescence temperature = 404°K * Results from total lineshape analysis. fin (V2?) 5.0 4.0 3.0 2.0 [.0 0.0 64 _- L. l l l I l 2.3 2.4 2.5 2.6 2.7 103/7°K Figure 13. Plot of ln(1/21) against 103/T°K for neat N,N-dimethylformamide-d1. 65 Table 15. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethylpropionamide. T°K 103/T°K 1* 1n(1/21) 5v* pA* sec Hz 314.2 3.183 0.0785 1.851 14.4 0.508 319.2 3.133 0.0455 2.396 14.5 0.504 321.0 3.115 0.0390 2.551 14.8 0.492 321.0 3.115 0.0393 2.545 14.5 0.504 322.2 3.103 0.0395 2.538 14.7 0.502 322.2 3.103 0.0358 2.640 14.8 0.507 323.5 3.091 0.0294 2.833 14.5 0.498 -323.6 3.090 0.0289 2.851 15.0 0.495 325.9 3.069 0.0268 2.928 14.7 0.502 325.9 3.069 0.0268 2.925 14.9 0.496 326.2 3.066 0.0249 2.300 327.5 3.053 0.0228 3.088 327.9 3.050 0.0209 3.176 329.3 3.036 0.0194 3.250 330.5 3.026 0.0172 3.369 332.9 3.004 0.0139 3.582 334.8 2.987 0.0118 3.750 337.2 2.966 0.0096 3.950 E8 = 18.9 i 0.4 kcal/mole logloA = 13.9 i 0.3 40*296 = 17.2 i 0.4 kcal/mole AH" = 18.2 1 0.4 kcal/mole 83* = 3.1 t 1.2 eu Coalescence temperature = 331°K * Results from total lineshape analysis. 66 4.20 - . 1' 3.80 3.40 an (yfi‘r', 3.00 2.60 2.20 1.80 I l l l l l 2.98 3.02 3.06 3.10 3.14 3.18 103/T°K Figure 14. Plot of 1n(1/21) against 103/T°K for neat N,N-dimethylpropionamide. Table 16. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat N,N-dimethyltrichloroacetamide. 67 T°K 283.2 286.6 288.0 290.1 292.1 295.0 296.8 298.2 303.6 306.9 103/T°I( 3.531 3.489 3.472 3.447 3.423 3.390 3.369 3.353 3.294 3.259 * T sec 0.0326 0.0233 0.0202 0.0171 0.0124 0.0106 0.0081 0.0074 0.0047 0.0033 1n(l/21) 1.729 3.067 3.207 3.378 3.695 3.851 4.119 4.212 4.666 5.030 * 5v Hz 29.3 29.2 29.1 29.2 29.0 29.2 29.1 0.498 0.496 0.504 0.512 0.502 0.495 0.501 Ea = 16.7 i 0.3 kcal/mole £67293 = 15.0 i 0.3 keel/mole AB* = 1.1 eu Coalescence temperature = 299°K * Results from total lineshape analysis. logloA = 14.0 i 0.2 2 [91 16.1 i 0.3 kcal/mole 90106:?) 5.40 5.00 4.60 4.20 3.80 3.40 3.00 2.60 I 3128 Figure 15. 68 L I I l l 3.32 3.36 ' 3.40 3.44 3.48 103/T °K Plot of 1n(1/21) against 103/T°K for neat N,N-dimethyltriChloroacetamide. 3.52 Table 17. for the internal rotation about the central C-N bond in neat N-ethyl-N—methylacetamide. Experimental data and calculated activation parameters T°K 325.7 328.7 331.8 333.1 334.8 337.8 338.2 340.1 340.8 344.2 347.2 349.1 103/T°K ' 3.070 3.042 3.014 3.002 2.987 2.960 2.957 2.940 2.934 2.905 2.880 2.865 31» T 88C 0.0685 0.0543 0.0347 0.0314 0.0250 0.0206 0.0193 0.0173 0.0169 0.0122 0.0092 0.0077 1n(1/21) 1.988 2.220 2.668 2.766 2.997 3.191 3.256 3.362 3.387 3.714 3.994 4.176 * 5v Hz 15.9 15.9 16.0 15.8 16.2 15.9 16.1 15.6 15.7 0.502 0.493 0.500 0.464 0.520 0.470 0.518 0.475 0.503 'Ea = 20.8 i 0.5 kcal/mole # AG 298 fist = 7.0 i 2.0 eu 18.0 i 0.5 kcal mole Coalescence temperature = 344°K * Results from total lineshape analysis. 10g1oA = 14.8 20.1 i 0.5 Real/mole l... .1 All...“ . in I'& 1’) 414C) 4‘”) 31K) 3120 2!“) 2A“) 21!) 70 I I l I I I I 2.86 2.90 2.94 2.98 3.02 3.06 3.10 103/T°K Figure 16. Plot of ln(1/21) against 103/T°K for neat N-ethyl—N-methylacetamide. I“ IF— 71 Table 18. Experimental data and calculated activation parameters for the internal rotation about the central C-N bond in neat Nfinfbutyl-N-methy1acetamide. T°K 103/T°K T* ln(1/21) 8v* pA* .222... Hz 329.0 3.039 0.0432 2.449 14.7 0.476 '333.8 2.996 0.0284 2.869 15.0 0.470 334.0 2.994 ‘ 0.0295 2.832 14.8 0.480 337.8 2.961 0.0190 3.268 15.1 0.474 338.2 2.957 0.0183 3.310 15.4 0.494 '341.1 2.931 0.0154 3.478 342.6 2.919 0.0134 3.618 345.1 2.898 0.0112 3.798 347.2 2.880 0.0089 4.030 347.8 2.875 0.0083 4.094 352.1 2.840 0.0060 4.421 163,0A a 14.1 2 0.3 20*... = 17.9 2 0.5 kcal/mole as? = 19.1 1 0.5 kcal/mole A13:t = 4.1 i 2.0 eu Ea = 19.7 2 0.5 Real/mole Coalescence temperature = 341°K * Results from total lineshape analysis. nnlfizlrl ‘16C1 4120> 31K) 33K) 3&X) 2!“) 21K) 72 0 CH3 ’0-111’ 1130’ ‘01121011212 0113 I I l I I 2134 Figure 17. 2.88 2.92 2.96 3.00 3.04 103/7°K Plot of 1n(l/2T) against 103/T°K for neat anfbutyl-N-methylacetamide. .1 “uh 'I W“. 57‘.” 1 73 lineshape analysis of the N-methyl proton resonance. The experimental data and the calculated activation parameters are tabulated in Table 19. The logarithm of the rate constant is plotted against 103/T°K in Figure 18. . N-Isopropyl-N-methylacetamide --- Hindered internal rotation in N- isopropyl-N-methylacetamide in the neat liquid was studied by total line- shape analysis of the N-methyl proton resonance. The experimental data f“‘ and calculated activation parameters are tabulated in Table 20. The logarithm of the rate constant is plotted against 103/T°K in Figure 19. Resonance Assignments in the Proton NMR Spectra of Several Tertiary Amides In this study, Eu(f0d)3, tris(l,l,l,2,2,3,3-heptaf1uor0-7,7-dimethy1- 4,6-octanedionato) europium (III), a paramagnetic shift reagent, was used to assign the proton resonances in substituted amides. Spectra were ob- tained at 100 MHz at a temperature chosen for each amide to be well below the coalescence temperature so that the rotational isomers are distin- guishable. Samples were 0.2 M amide in C014 and increasing amounts of Eu(fod)3 were added. The chemical shifts relative to TMS were plotted versus the mole ratio of shift reagent to amide. A linear correlation was found for each group of protons. A linear least-squares analysis was applied to each data set to ascertain the chemical shift in the absence of n=1 n=0 8 ) value for each group of shift reagent and the AEu (AEu = SCC14 ' C014 protons. The chemical shift in the absence of shift reagent was used to assign the resonance to the corresponding group of protons. l-Methyl-Z-pyrrolidinone --- To establish the validity of the method, 1-methy1-2-pyrr01idinone was chosen as a model system since the Namethyl 74 Table 19. EXperimental data and calculated activation parameters for the internal rotation about the central C-N bond in . neat N-cyclohexyl-Ndmethylacetamide. 1°x 103/1°x 1* 1n(1/21) 6v* pA* .2251._ Hz 305.1 3.277 0.1439 1.245 11.1 0.457 314.1_ 3.184 0.0506 2.290 11.2 0.447 318.5 3.140 0.0364 2.621 11.1 0.470 320.7 3.118 0.0347 2.667 10.7 0.487 321.8 3.107 0.0304 2.801 10.7 0.485 322.3 3.102 0.0292 2.840 11.1 0.450 324.7 3.080 0.0212 3.159 326.8 3.060 0.0171 3.376 330.1 3.029 0.0124 3.699 'Ea'= 18.9 1 0.5 kcal/mole logloA = 14.1 1 0.3 20*296 = 17.1 i 0.5 kcal/mole 0H3 = 18.3 1 0.5 kcal/mole AS: = 3.9 i 2.0 cu Coalescence temperature = 325°K * Results from total lineshape analysis. 4420 3!“) 3040 3110 1.8CI L4!) IIX) 75 I I I I I I I 3.02 3.06 3.10 3.14 3.18 3.22 3.26 103/7°K Figure 18. Plot of 1n(1/21) against 103/T°K for neat Necyclohexyl—N-methylacetamide. Table 20. for the internal rotation about the central C-N bond in neat N-isopropyl-N-methylacetamide. 76 Experimental data and calculated activation parameters T°K 305.6 314.8 316.7 318.9 320.8 322.3 325.4 326.6 330.8 103/T°K 3.272 3.177 3.157 3.136 3.117 3.102 3.073 3.062 3.023 * T 88C 0.1028 0.0453 0.0414 0.0335 0.0278 0.0233 0.0181 0.0156 0.0104 1n(1/21) 1.581 2.402 2.492 2.704 2.888 3.066 3.319 3.467 3.873 * 8v Hz 11.4 11.5 11.2 11.2 11.3 11.2 0.444 0.459 0.441 0.457 0.440 0.448 E = 18.1 1 0.5 Real/mole ...: \I 0 1+ 2.0 eu 0.5 kcal/mole Coalescence temperature = 322°K * Results from total lineshape analysis. log1oA = 13.6 i 0.3 Aflt 17.5 t 0.5 kcal/mole 91111511 4.20 3. 80 3u40 3.00 2.60 2.20 1u80 |.40 77 _ 0 0113 §C%NI 1130’ \c111c1-1312 I I I I I I. I 3.02 3.06 3.10 3.14 311 8 3 22 3.26 1/1'°1< x103 Figure 19. Plot of ln(1/21) against 103/T°K for neat N-isopropyl-N—methylacetamide. 78 group is fixed in the gig (to oxygen) position. The proton resonance of the N-methyl group can be identified even without the addition of the shift reagent because the resonance appears as a singlet, the only sing- let in the spectrum. The chemical shifts from the addition of Eu(fod)3, and the results of the least-squares analysis, are given in Table 21. As expected the gig-methyl protons were shifted to a greater extent than the Eggggémethylene protons in solutions containing the amide complexed with the shift reagent. The AEu value for the gigfmethyl protons is 10.12 ppm and for the Egggggmethylene protons is 5.49 ppm. For the re- maining amides the resonances of the gigfggagg pair which are shifted the greater amount are assigned to the gig group. N,N-Dimethylformamide --- The proton spectrum of N,N-dimethylforma- mide shows three resonances. The formyl proton resonance appears at 7.86 ppm and the two N-methyl protons resonate at 2.93 and 2.81 ppm. The re- sults from the addition of Eu(f0d)3, and the least-squares analysis, are tabulated in Table 22. For the uncomplexed amide the chemical shift at 2.81 ppm was assigned to the gisfmethyl protons (AEu = 9.39 ppm) and the chemical shift at 2.93 ppm to the trans-methyl protons (AEu = 4.04 ppm). N,N-Diethylformamide --- The proton NMR spectrum of N,N-diethyl- formamide shows five resonances: a singlet at 7.90 ppm from the formyl proton, two quartets at 3.30 and 3.26 ppm from the N-methylene protons, and two triplets at 1.19 and 1.10 ppm from the methyl protons. The data from the addition of Eu(f0d)3 are tabulated in the Table 23 along with the results of the least-squares analysis. The quartet at 3.30 ppm was assigned to the gig (to oxygen)-methy1ene proton (AEu = 9.86 ppm) and the quartet at 3.26 ppm to the-Egang-methylene protons (AEu = 3.87 ppm). The triplet at 1.19 ppm was assigned to the trans-methyl protons (AEu = 2.48 79 Table 21. Proton chemical shiftsa, observed and calculated, in lemethyl-Z-pyrrolidinone, with increasing amounts of Eu(fod)3. 0 §§§ CHa(1) C--N /’ (4)H2C CHz(2) \/\ ‘\\ B N A $3 Mole Ratio of Shift Reagent/Amide 0.00 0.097 0.194 0.291 0.00 8 b 8 Proton Resonance obs calc AEu (1) 2.76 3.78 4.78 5.75 2.80 i 0.02 10.12 i .08 (2) 3.32 3.87 4.41 4.93 3.34 i 0.01 5.49 i 0.05 (3) 1.98 2.49 2.99 3.48 2.00 i 0.01 5.09 i 0.05 (4) 2.19 3.54 4.85 6.12 2.25 i 0.03 13.33 i 0.12 aSpectra'were obtained at 32°C. bA11 chemical shifts are in ppm from TMS. 80 Table 22. Proton chemical shiftsa, observed and calculated, in N,N-dimethylformamide with increasing amounts of Eu(fod)3. 0 CH3(1) / , /C—N\ (3)H CH3(2) Mole Ratio of Shift Reagent/Amide 0.00 0.084 0.166 0.248 0.330 0.00 0 b 6 Proton Resonance obs calc AEu (1) 2.81 3.65 4.37 5.18 5.97 2.85 i 0.03 9.32 i 0.16 (2) 2.93 3.30 3.60 3.96 430.1 2.95 i 0.02 4.04 i 0.10 1 , (3) 7.86 9.09 10.12 11.29 12.43 7.94 i 0.06 13.42 i 0.20 [—7 aSpectrawere obtained at 32°C. bAll chemical shifts are in ppm from TMS. 81 Table 23. Proton chemical shiftsa, observed and calculated, in N,N—diethylformamide with increasing amounts of Eu(fod)3. 0 CH2(1)-C33(2) \ / C'——N 1’ ‘\ (5)H CHz(3)-CH3(4) Mole Ratio of Shift Reagent/Amide 0.00 0.100 0.200 0.300 0.400 0.0 b Proton Resonance 6Obs 5calc AEu (1) 3.30 4.33 5.40 6.34 7.24 3.35 i 0.07 9.68 i 0.27 (2) 1.10 1.72 2.35 2.92 3.46 1.12 i 0.04 5.79 i 0.15 (3) 3.26 3.68 4.11 4.49 4.84 3.26 i 0.03 3.87 i 0.12 (4) 1.19 1.46 1.74 1.98 2.21 1.18 i 0.02 2.48 i 0.08 (5) 7.90 9.29 10.23 12.01 13.23 7.99 i 0.10 13.10 t 0.36 aSpectrawere obtained at 32°C. bA11 chemical shifts are in ppm from TMS. 82 . ppm) and the triplet at 1.10 ppm to the cis-methyl protons (AEu = 5.79 9991) . §,N-Diisopr0py1formamide --- The proton NMR spectrum of N,N-diiso- propylformamide shows five resonances: the formyl proton resonates at 8.03 ppm, the N-methine protons resonate at 3.87 and 3.55 ppm, and the methyl protons resonate at 1.27 and 1.26 ppm. The data from.the addition of Eu(fod)3 in Table 24 along with the results of the least-squares anal- ysis. The chemical shift at 3.87 ppm was assigned to the gig (to oxygen) methine proton (AEu = 8.24 ppm) and the chemical shift at 3.55 ppm was assigned to the Eragggmethine proton (AEu = 3.07 ppm). The assignment for the methyl protons is very uncertain because of the proximity of the resonances. Tentatively, the resonance at 1.27 ppm was assigned to the Eggggrmethyl protons (AEu = 2.14 ppno and the resonance at 1.26 ppm to the gigfmethyl proton (AEu = 6.03 ppm). N,N-Dimethylacetamide --- The proton NMR spectrum.of N,N-dimethyl- acetamide shows three resonances. The N-methyl protons resonate at 2.99 and 2.85 ppm and the acetyl protons resonate at 1.96 ppm. This amide was studied by use of two different shift reagents, Eu(fod)3 and tris(l,l,l,- .2,2,3,3-heptaf1uoro-7,7-dimethyl-4,6-octanedionato) praseodymium (III), Pr(fod)3. The shift reagent Pr(f0d)3 produces upfield shifts, however the resultant effect is identical to Eu(fod)3. The results from the least- squares analysis and the addition of Eu(fod)3 are given in Table 25 and the results from Pr(fod)3 in Table 26. The resonance at 2.99 ppm was assigned to the trans-methyl protons (AEu = 5.24 ppm, APr = -9.04 ppno and the resonance at 2.86 ppm to the gig-methyl protons (AEu = 10.18 ppm, APr = -16.16 ppm). 83 Table 24. Proton chemical shiftsa, observed and calculated, in_ N,N-diisoprOpylformamide with increasing amounts of Eu(fod)3. 0 CH(1)-(Cfla)2(2) ‘§§(:-11/’ \ (518/ cn<3>- Mole Ratio of Shift Reagent/Amide 0.00 0.100 0.200 0.300 0.400 0.00 I 5 b 0 Proton Resonance Obs calc AEu (1) 3.87 4.80 5.60 7.27 8.28 3.96 i 0.02 8.24 i 0.10 (2) 1.26 1.90 2.46 3.07 3.70 1.27 i 0.03 6.03 i 0.10 4 Y! F" (3) 3.55 3.95 4.24 4.55 4.87 3.62 1 0.01 3.07 1 0.05 "' (4) 1.27 1.50 1.70 1.93 2.14 1.28 i 0.01 2.14 i 0.03 (5) 8.03 9.18 10.21 11.29 12.42 8.07 i 0.04 10.81 i 0.16 aSpectrawere obtained at 32°C. bAll chemical shifts are in ppm from TMS. 84 Table 25. Proton chemical shiftsa, Observed and calculated, in N,N-dimethylacetamide with increasing amounts of Eu(f0d)3. 0 'CH 1 // 3( ) \\ (3)1130/ 0113(2) Mole Ratio of Shift Reagent/Amide 0.00 0.100 0.200 0.300 1555; 5-77’ Proton Resonance 6obsb 5calc AEu (1) 2.85 3.92 4.94 5.96 2.90 1 0.03 10.18 1 0.02 (2) 2.99 3.50 4.02 4.54 2.97 1 0.005 5.24 1 0.01 (3) 1.96 3.17 4.30 5.45 2.03 1 0.01 11.39 1 0.06 5:75 aSpectra‘were obtained at 32°C. bA11 Chemical shifts are in ppm from TMS. 85 Table 26. Proton chemical shiftsa, observed and calculated, in N,N-dimethylacetamide with increasing amounts of Pr(fod)3. 0 CH 1 //, 3( ) (3183/ \011312) Mole Ratio of Shift Reagent/Amide 0.00 0.095 0.190 0.285 5 Proton Resonance obs (1) 2.85 1.32 -0.16 -1.74 (2) 2.99 2.09 1.29 0.38 (3) 1.96 0.34 -l.17 -2.92 0.00 5calc 2.87 i 0.03 2.97 i 0.02 H- 2.02 .05 APr -l7.22 i 0.10 0.09 0.16 aSpectrawere obtained at 32°C. bA11 Chemical shifts are in ppm from TMS. 86 N,N-Dimethylcarbamylch10ride —-- The proton spectrum of N,N-dimethyl- carbamylchloride shows two resonances,at 3.17 and 3.06 ppm,for the N- methyl protons. The results from the addition of Eu(fod)3 and the least- squares analysis are given in Table 27. The chemical shift at 3.17 ppm was assigned to the Eggggfmethyl protons (AEu = 6.18 ppm) and the chemie cal shift at 3.06 ppm to the gigfmethyl protons (AEu = 11.65 ppm). N,N-Dimethyltrichloroacetamide --- The proton spectrum of N,N- dimethyltrichloroacetamide shows two resonances,at 3.38 and 3.10 ppm,from the N-methyl protons. The results from the addition of Eu(fod)3 and the least-squares analysis are tabulated in Table 28. The resonance at 3.38 was assigned to the Eggngimethyl protons (AEu = 7.99 ppm) and the reso- nance at 3.10 ppm to the gigfmethyl protons (AEu = 16.63 ppm). N,N-Dimethyltrifluoroacetamide --- The proton NMR spectrum of N,N- dimethyltrifluoroacetamide shows two resonances, at 3.13 and 3.02 ppm, from the N-methyl protons. Each of these resonances is split into quar- tets due to the long-range coupling with fluorine (I = 1/2). The results from the addition of Eu(f0d)3 and the least-squares analysis are given in Table 29. The chemical shift at 3.13 ppm was assigned to the trans- methyl protons (AEu = 5.12 ppm) and the chemical shift at 3.02 ppm to the gigfmethyl protons (AEu = 9.38 ppm). Ethyl-N,N-dimethy1carbamate --- The proton NMR spectrum of ethyl- N,N-dimethy1carbamate shows four resonances. The methylene and methyl protons of the ethoxy group resonate at 4.18 and 1.32 ppm, respectively. The N-methyl protons resonate at 2.94 and 2.92 ppm. The results from the addition of Eu(fod)3 are given in Table 30. The resonances assignments for the N-methyl protons could not be made with certainty because the chemical shift (2.0 Hz) between the cis and trans resonances is smaller 87 Table 27. Proton chemical shiftsa, observed and calculated, in N,N-dimethylcarbamylchloride, with increasing amounts of Eu(fod)3. O /CH3(1) ‘ / \ Cl 033(2) MOle Ratio of Shift Reagent/Amide 0.00 0.099 0.198 0.297 0.00 8 b 8 Proton Resonance obs calc afiu (1) 3.06 4.10 5.37 6.43 2.97 i 0.08 11.65 i 0.30 (2) 3.17 3.71 4.38 4.94 3.11 i 0.04 6.18 i 0.10 aSpectra‘were obtained at -22°C. bAll chemical shifts are in ppm from TMS. 88 Table 28. Proton chemical shiftsa, observed and calculated, in N,N—dimethyltrichloroacetamide with increasing amounts of Eu(fod)3. 0 /CH3(1) \/ 013C \C'H3(2) Mole Ratio of Shift Reagent/Amide 0.00 0.097 0.194 0.291 0.00 8 b 8 Proton Resonance obs calc AEu (l) 3.10 4.75 6.40 7.92 3.18 i 0.07 16.63 i 0.30 (2) 3.38 4.16 4.92 5.68 3.40 i 0.002 7.99 i 0.02 aSpectra‘were obtained at -22°C. bAll chemical shifts are in ppm from TMS. 89 Table 29. Proton chemical shiftsa, observed and calculated, in N,Nedimethyltrifluoroacetamide‘with increasing amounts of Eu(fod)3. 0 /CH3(1) // '\\ F3C/ 013(2) Mole Ratio of Shift Reagent/Amide 0.00 0.100 0.200 0.300 0.00 8 b 8 Proton Resonance obs calc (l) 3.02 3.97 4.92 5.94 2.97 i 0.04 (2) 3.13 3.58 4.09 4.60 3.07 i 0.005 9.38 5.12 AEu i 0.14 i 0.02 8Spectra were obtained at 32°C. bAll chemical shifts are in ppm from TMS. 90 Table 30. Proton chemical shiftsa, observed and calculated, in ethyl-N,N-dimethylcarbamate with increasing amounts of Eu(fod)3. 0 ‘ CH 1 ¢§§ ” 3( ) C--N / \ (4)H30-(3)H2C0 CH3(2) Mole Ratio of Shift Reagent/Amide 5537 0.00 0.100 0.200 0.300 0.00 b Proton Resonance aobs .. 5calc AEu (1) 2.92 3.89 4.88 5.75 2.97 i 0.06 9.80 i 0.21 (2) 2.94 3.48 4.11 4.66 2.90 i 0.05 6.24 i 0.16 ‘psj (3) 4.18 5.65 7.12 8.44 4.28 i 0.09 14.55 i 0.30 (4) 1.32 1.68 2.07 2.48 1.28 i 0.01 4.18 i 0.05 aSpectrawere run at -22°C. bAll Chemical shifts are in ppm from TMS. 91 than the error of the analysis. The AEu value for the gi§.(to carbonyl oxygen) methyl protons is 9.80 ppm and for the Eggggfmethyl protons is 6.24 ppm. N-Isopropyl-N-methylformamide --- The proton NMR resonance of N- Isopropyl-N-methylformamide shows eight resonances, four from each rota- tional isomer. The formyl proton chemical shift is different for each isomer being 7.97 and 7.84 ppm. The N-methyl protons resonate at 2.78 and 2.67 ppm. The N-methine protons resonate at 4.53 and 3.78 ppm. The methyl protons of the isopropyl groups resonate at 1.23 and 1.13 ppm. The re- sults from the addition of Eu(fod)3 and the least-squares analysis are given in Table 31. The resonance at 2.78 ppm.was assigned to the trans (to oxygen) N-methyl protons (AEu = 4.20 ppm) and the resonance at 2.67 ppm to the gi§_N1methyl protons (AEu = 9.81 ppm). The resonance at 4.53 ppm was assigned to the gi§_(to oxygen) N-methine proton (AEu = 14.14 ppm) and the resonance at 3.78 ppm to the trans N-methine proton (AEu = 4.14 ppm). The resonance at 1.23 ppm was assigned to the trans: methyl proton (AEu = 2.62 ppm) of the isopropyl group and the resonance at 1.13 ppm, to the gigfmethyl protons (AEu = 4.86 ppm) of the is0propy1 group. Carbon-13 Chemical Shifts of Aliphatic Amides The carbon-l3 chemical shifts given in this section were obtained at .a radiofrequency of 25.1 MHz on a Varian HA-lOO NMR spectrometer with n0:1se-modulated proton decoupling. The chemiCal shifts were initially rEtferenced to an external sample of 57% enriched carbon-13 methanol and Converted to the TMS standard by the following relationship: 92 Table 31. Proton chemical shiftsa, observed and calculated, in N-is0propyl-N-methylformamide with increasing amounts of Eu(fod)3. 0\ /CH3(1) 0\ /CH(5)-(CH3)2(6) /C——N\ c—b / C—N\ (4)11 CH(2)-(CHa)2(3) (8)11 0113(7) Mole Ratio of Shift Reagent/Amide 0.00 0.096 0.193 0.289 _Q__(_)_0_ ’ 5 b 8 Proton Resonance obs calc AEu (1) 2.67 3.60 4.57 5.46 2.68 i 0.05 9.81 i 0.21 (2) 3.78 4.16 4.57 4.95 3.77 i 0.02 4.14 i 0.09 (3) 1.23 1.49 1.74 1.99 1.24 i 0.003 2.62 i 0.01 (4) 7.97 9.26 10.64 11.89 7.96 i 0.08 13.82 i 0.35 (5) 4.53 5.76 7.15 8.45 4.43 i 0.06 14.14 i 0.25 (6) 1.13 1.56 2.02 2.49 1.10 i 0.001 4.86 i 0.01 (7) 2.78 3.14 3.56 3.94 2.74 i 0.03 4.20 i 0.15 (8) 7.84 9.06 10.41 11.68 7.76 i 0.05 13.77 i 0.25 aSpectra‘were run at 32°C. bAll chemical shifts are in ppm from TMS. 93 8mg = 50113011 + 49.4 (42) In order to comply with the accepted convention for proton NMR spectra, carbon-13 chemical shifts downfield from TMS are positive and upfield shifts from.TMS are negative. The uncertainty in the carbon-13 chemical shift measurements is i 0.1 ppm. Time-averaging with a Varian C-1024 computer was employed in some E——7. cases to increase the signal-to-noise ratio; however, as a result of the ; rolling baseline discussed in the Experimental Section, the number of i scans possible was limited. Fortunately, most of the resonances were i 4 detected after a single scan. Some of the carbon-13 spectra were obtained Ewe) below the ambient temperature so that both rotational isomers could be distinguished. Chemical shift assignments for the first several amides in each series were rather simple to make. Using these systems as a base, along with chemical shift data available in the literature, assignments were then made for the rest of the amides. Monosubstituted Amides The carbon-13 chemical shifts for fifteen monosubstituted amides are given in Table 32. The proton NMR spectra of monosubstituted formamides exhibit two sets of resonances, one from the conformer for which the N- alkyl substituent is gig (to carbonyl oxygen) and one from the correspond- ing trans conformer (114). However, one configuration was dominant; its fraction decreased from 0.92 for N-methylformamide to 0.88 for N-ethyl- formamide and N-isopropylformamide, and to 0.82 for N-tfbutylformamide. 94 .4280 as 8640526. some . a .u .8 @3988 we coouoo T909280 no somehow: 3 02/330» nowufimod conumoo .mzH Eoum Ema CH mum nonfizm ~00H50noo .o.mm “4 68856316 6063 608661. .246 m.NN n.0m o.mm m.- m.oN n.mm 3.0H w.a~ w.- n.m~ «.mm o.oN ¢.mm m.oH o.m~ m.NN 01m 018 uoosuwumosm Hzcooumo 3.8m H.2m 4.282 86384206821351-mrz H.3H m.oN m.~m 6.8m 6.05. muaamumomasuammm-z 0.32 8.0N H.Nm N.wm 8.N82 mousseuooasuam-a-z m.m~ n.23 1.8na umuaamususeomaasaouaOmH-z m.m~ n.23 N.Oea sesamuoumasaouaomH-z w.- n.03 n.262 66388246021660568H-z m.m2 1.3m 3.4na 66628441355684H2232-2 o.m2 5.3m H.mea 66024864168121231-2 1.32 4.3m 3.Hea 6642486062120m12 1.32 3.mm m.~oe 06224846021208-z 3.4m 1.14. 68688646535212362-2 «.mm 5.142 66.2441356645212062-z H.8N o.me~ 68388864a68a212302-z H.8N «.mea 684880004212382-z n.3m w.m82 measmauooaseuuz-z 8-8 o-N o-u 60-8 no meeea ucosuwumnam oowouuez .868388 663533346566868-z use. :4 6.4830426 46648820 me-coeumo .mm afiema 95 In all other monosubstituted amides, only one resonance was observed in the proton NMR spectrum for each group of protons of the N-alkyl sub- stituent. The carbon-13 NMR spectrum of each monosubstituted amide exhibits only one set of resonances for the N-alkyl substituent. The spectrom- eter was not able to detect the less abundant configuration in the formamides. In N-monosubstituted amides, the preferred configuration has been shown to be the one in which the N-alkyl substituent is gig to carbonyl oxygen by a variety of methods including dipole-moment measurements and infrared, Raman, ultraviolet spectroscopy (for a detailed bibliography see reference 114). Thus, the carbon-13 resonances of the N-alkyl sub- stituents in Table 32 are for the configuration in which the substituent on nitrogen is cis to carbonyl oxygen. Symmetrically Disubstituted Amides The carbon-l3 chemical shifts for twenty-four symmetrically N,N- disubstituted amides are given in Table 33 (N,N-dimethylamides) and Table 34 (N,N-diethylamides, N,N-diisopropylamides, and N,N-di-nfprOpylamides). In some cases, the spectrum of a particular amide was recorded at tempera- ture of 0°C 80 that both rotational isomers could be distinguished. The chemical shift assignment of the N-methyl substituents in di- methylamides was based on an expected upfield steric shift for the methyl group gig to the carbonyl oxygen. This expectation was confirmed experi- mentally for N,N-dimethylformamide, N,N-dimethylacetamide, and N,N- dimethylcarbamylchlorida by McFarlane (55), who found that the upfield 96 . . .. 8 .n .8 008.0800 88 8048.808 8.303.800 .80 880.8880 3 0830.808 8808080048 Conumou .Uoo um 88088.88...qu 0.803 0.380480 .8050 880 .93 80 80580qu 0.803 0.88045 n .988. 80.3 885 888 0.80 08.8330 8008.880an 8.88 8.88 8.88 8.88 8.888 888888888888888881z.2184288 8.448 8.888 4.88 8.48 8.888 8888884888888888881z.z 4.88 8.88 8.88 8.888 888888888686888888888888812.z 8.88 8.48 4.88 8.888 88828888868682888888888881z.z 8.88 8.88 8.88 8.488 0888888886868288428888812.z 8.88 8.88 8.888 88886888848888888488088812.z 8.88 8.88 8.848 68888686588888842888881z.z 8.88 8.88 8.88 8.88 8.848 888888858884884888uz.z 8.88 8.88 8.88 8.48 8.848 8888888888688848882881z.z 8.88 8.88 8.88 8.88 8.48 8.848 8888888858Lmr848808881z.z 8.8 8.88 8.88 8.48 8.848 80888888886888488088812.z 8.84 8.88 8.88 8.848 88888888888428888812.z 8.88 8.88 8.888 86888888688488888812.z IWHNI .Ioflwl Iflwol 888 “80.8.8181 018 0.1.0 088888.84 unguaumnfim us0ouwuwn=m 8.889.800 a0wouuaz .808888882808881z.z 0868 88 8880828 88888888 881868888 .88 08884 97 o o o o xfi an ad vmuocmfi m.“ 000800 88000800 80 00008080 00 0>800808 00808000 0008000 .080 00 00080000 0803 0800000 880 n .mZB 808m 800 08 080 000800 800800000 8.88 8.88 0.88 0.88 0.m~ 8.88 8.0m 8.88 8.88 8.88 0.88 8.88 8.88 0.08 0.88 0.88 m.m8 8.08 8.08 8.88 8.88 w.m8 8.0m 8.08 «.mw 0.88 0.88 0.888 8.88 m.m8 8.88 8.88 m.m8 8.M8 8.m8 8.88 8.88 0.88 ~.mm w.m8 8.m8 8.08 8.88 0.08 0.88 o.m8 0.08 8.88 0.88 8.88 8.88 m.oq n.88 8.m8 «.ma 8.8m n.88 00080 080 080 00080 080 00080 0-4 0-0 0-4 0-4 0-8 0-8 0-0 0-8 0000080000m 00000800000 88000800 0000808z 8.00H o.moH w.Hna o.m08 o.~oH N.moH 8.008 n.88H «.mma 0.08H 0.Noa o 0. 008800000880080Lmlanlz.z 008008808880080Lmlwalz.z 008000080080880080008Hntz.z 00880000088008000880Iz.z 0080088OMH800800088QIZ.Z 00800888008800080Iz.z 008000000080800885008012.z 0080088000LMI8800080IZ.2 0080000800808800080|z.z 008000000880008nlz.z 0080058088800080Iz.z 88888 .000830 0000080000080Iz.z 8880088000350 80000 0000 80 8800888 48880888 88-888888 .88 84888 98 carbon-l3 shift of the N-methyl substituents in each of these amides was .Eifi to the carbonyl oxygen. The carbon assignments in N-alkyl substituents larger than methyl were made with the aid of the two lanthanide shift reagents, Eu(fod)3 and Pr(fod)3. The effect of the shift reagent on an amide will be the same in the carbon-13 spectrum as in the proton spectrum; that is, the resonance of a gigfitrgng pair which is shifted the greater amount can be Eff“ assigned to the group gig to carbonyl oxygen. The results of the addi- tion of 0.2 grams of Eu(fod)3 to a five percent by volume solution of N,N-dijn-propylformamide in carbon tetrachloride and 0.2 grams of Pr(fod)3 T to a five percent by volume solution of N,N-di-nfprOpylformamide in carbon L ,J tetrachloride are given in Table 35. Eu(fod)3 shifts the resonances downfield and Pr(fod)3 shifts the resonances upfield. For the EEETEEEEQ pair<1 to nitrogen, the carbon resonance upfield was shifted to greater amount and was assigned to the carbon gig to carbonyl oxygen . The same result was found for the gigfitggns pair 6 to nitrogen; however, for the ‘gig-tgang pair 7 to nitrogen, the downfield resonance was shifted the greater amount and was assigned to the carbon in the substituent gig to carbonyl oxygen. The assignments of carbon resonances for the N-alkyl substituents in the carbon-l3 spectra of the rest of the amides are based on this experiment. The carbon-l3 spectrum of N,N-dimethylpivalamide shows a single reso- nance for the N-methyl carbons. The reason for this is that the coales- cence temperature is below the freezing point of this amide. 99 .ix “8 %1 1H}, .8 ”8|! I1]. 1.1!.1f1L .029 5080 500 08 080 800800 400800000 0 o o o o o 0 ma 8 88 8 84 8 88 8 88 8 88 8 88 4 48808888 8 . . . . . . . 029 0 84 0 N8 8 88 0 mm 8 00 N 00 AmAwowvsmv 0 . . . . . . 0:9 0 04 N 88 0 ON 0 48 m 08 m 08 40000000 00800 ozv 0 00080 080 880 80080 080 00080 8 48808880888888 8 48888880888082 8 48880888888082 . .0088040008000 008800 08 008005800480080Lmn80uz.z mo 000000808800 48380-2 000 800 0000800 80088000 08-008800 no 0005008800 .00 08809 100 Unsymmetrically Disubstituted Amides The carbon-l3 chemical shifts for eleven unsymmetrically N,N—disub- stituted amides are given in Table 36. All of the Spectra were obtained at 0°C so that both rotational isomers could be distinguished. The assignment of the N-alkyl substituents are based on the results of the effect of the lanthanide shift reagent on N,N-di-gfpropylformamide. The carbon-13 NMR spectra of N-ethyl-N-methylpivalamide and N‘Ef butyl-N-methylpivalamide exhibit only one set of resonances for each set of N-alkyl substituents. The reason for this observation is that the coalescence temperature is below the freezing point of each of these amides. The carbon-13 spectra of Netfbutyl-N-methylformamide and N-E; butyl-N-methylacetamide also exhibit only one set of resonances for each N-alkyl substituent. .This may be a result of the coalescence temperature being below the freezing point in these amides or one of the rotational isomers may be strongly preferred over the other (49). 101 . ... A .0 .H00000000 08 000800 Hm0o0800 80 00008080 00 0>800H08 00808000 0008000 0 .mzH 808m 800 08 080 00mw00 H00H50000 m.w~ m.m~ 0.00 0.00 ~.mm ~.m~ H.mm H.mm 0.0a 0.0H m.o~ m.o~ 0.0m 0.0m 0.00 0.00 0.0a 0.0a m.o~ w.o~ n.0m H.Nm 5.50 m.mq 0.0a 0.0H w.o~ N.HN m.om 0.Hm ~.n¢ 5.00 0.0H 0.0H N.o~ o.o~ H.0N m.om w.mq m.wq m.ma w.o~ m.mq «.mq m.mH N.HN m.~q 8.00 m.ma m.mH N.mq ~.mq m.NH 0.0H m.Nq 0.00 m.NH m.qH H.0m m.¢0 080 .mmmmm. mmmmmw 080 080 00080 080 Mmmmmw 0-0. 0:0 Hmmamsz m.mm m.mm m.Hm m.Hm n.0m 8.0m 0.mm 0.0m m.~m H.0m m.m~ 0.0m 0.0m N.mN 0.0N m.w~ 0.0m 0.0m «.mm 0.mm m.wm 0.0m 080 mmmmmw Awwnumauz 000000800000 00w080wz «88080008880080-2-0musmnmrz 088080808880000-2-88058L0rz 00880H0>HQH00008IZIHmusmLmnz 08808880580888800meuzna8uamumrz 00H500000H00008IZIH0000Lmrz 88280808885800-1380; 008000000Hh00maIZIH>008000le 008808800H>000012ia>008000le 0880888>888800mauzu88numuz m88080808888080-zu88008-z 888amEHomamsumauzu88au8-2 08808 .000800 0000080000080|z.z 0HH008800000000 0000 m0 0.0000800 H00H0000 mslccnumo .cfi ”Hana 102 8.0m 0.0m 0.0m DIN— 0.08 0.00 8.00 0.8m 0.08 0.00 0.00 0nd 000000800000 80000800 5.058 ~.808 0.058 0.058 0.008 0.008 0.008 0.808 0.058 0.058 5.008 00800000080000E|zn800smhwrz 888080808Hmnumenzu88uamumuz 0880888>80880080-2-88028Lmrz 0088080000008800000IZI80000Lmuz 00800000080000EIZI80000Lmrz 008008800800000IZI80000Lmlz 00800000080000BIZI800080008IZ 008808800800008|ZI80008000812 0880888>8088gumenzu88008-2 m8808000888numenzn88sumuz 08808080888sumauzu88000uz 08808 8.0.00000 .00 08009 DISCUSSION Hindered Internal Rotation Total Lineshape Analysis The only reliable method for extracting kinetic results from NMR { data is by direct comparison of the experimental lineshape with the spec- i ,9 trum calculated using the theoretical lineshape equation. To obtain the best values of the NMR parameters a curve fitting procedure is required and this can best be achieved by means of the digital computer. For ex- change between two sites, the parameters required are: (a) the chemical shift in the absence of exchange, 6v; (b) the relative populations at sites A and B, pA and pB; (c) the spin-spin relaxation times in the absence of exchange, l/T'2A and l/TéB, as measured from the individual linewidths; and (d) the rate constant for the exchange process 1/21. The chemical shift in the absence of exchange was initially deter- mined at a temperature where the exchange process is slow. It was then introduced. into the lineshape analysis as an adjustable parameter to be determined at each temperature below the coalescence temperature. For the amides studied in this research the calculated chemical shift in the ab- sence of exchange varied unsystematically about the value initially mea- sured. However, it has been shown (24) that the chemical shifts of each resonance do change individually with temperature relative to some 103 104 standard, but the chemical shift difference between the two resonances remains fairly constant over a large temperature range. The relative populations, and p8, should be exactly 0.5 for PA symmetrically disubstituted amides. For the unsymmetrical amides pA may vary between zero and unity, but should approach 0.5 with increase in temperature. In practice, this factor changes only slightly with temp perature from the value measured in the absence of exchange. Gutowsky f—T‘ .gg _l. (29) were able to use the constant value pA = 0.54 over the entire temperature range covered in the investigation of the barrier of N-methyl- N-benzylformamide without introducing appreciable error into the rate constants. In the research described in this thesis, the curve fitting program calculated values of pA and pB at each temperature up to the coalescence temperature. The spin-spin relaxation times as measured from the linewidths in the absence of exchange become relatively unimportant as the temperature increases and exchange broadens the linewidths. These values have no particular significance since they include inhomogeneity broadening, and the rate constants are not appreciably altered if these values are held constant throughout the entire temperature range studied. The Frequency Factor in the Arrhenius Equation The question of what frequency factors might reasonably be expected in unimolecular reactions is discussed at length in Glasstone, Laidler and Eyring (115). According to absolute reaction rate theory the normal expectation is for the frequency factor to lie in the range 1013 to 1014 sec'1 for the gas phase; Kondrat'ev (116) has reached the same conclusion. 105 Drastic exceptions can be expected only for rather unusual molecules such as those whose activated state may be a triplet electronic state with the transition back to the singlet ground state highly forbidden. Smaller departures in the liquid phase could result if the molecular ground and excited states were quite dissimilar in some important aspect such as interaction with solvent molecules. However, with amides, even if there is preferential solvation of one state or the other, it seems unlikely that the preference should change greatly from amide to amide (16). Thus, it seems reasonable that amide frequency factors should lie in the range 1013 to 1014 sec'l. Rotational Barriers in N,N-Disubstituted Amides The energy barriers for hindered internal rotation about the C-N bond in N,N—disubstituted amides is believed to result largely from the partial double-bond character of this bond. There is a variety of evi- dence that the resonance structures of Figure I contribute nearly equally to the resonante hybrid (117,118). However, there is an insufficient amount of reliable data available to evaluate the substituent effects of R1, R2, R3 on the rotational barrier, despite the many articles that have appeared on the subject. Thus, an attempt has been made in this research to obtain more precise values for a series of N,N-dimethylamides and to extend the method of analysis to four unsymmetrically N,N-disubstituted amides. The results are summarized in Tables 37 and 38. The barrier heights in amides are indicative of the degree of stabil- ization of the approximately planar ground state relative to the non- planar activated state. The factors that tend to stabilize the planar as! L... 0,- .08008000 000000088 80000 08 0080080000 003 000080 805808 00 00880000 00008-00040 .08008000 000000088 80000 08 000008000 003 000080 808800 00 00880000 00008-00080 0803 8808 8.888.8 8.088.88 8.088.88 8.088.88 8.088.88 020 888 888.8- 8.080.88 8.088.88 - 8.088.88 0:0 88 8.88 8.88 8.88 88 0.88 8.8888 88.85888888820858928 x803 8808 8.888.8 8.088.88 8.080088 8.080.88 8.088.88 80820 88 8.088.o 8.088.88 8.080.88 8.088.88 8808208 88808088808080088088008080-z.z x803 8808 8.880.8 8.080.88 8.088.88 0.088.88 8.088.88 0020 08 0.888.8- 8.88 0.088.88 8.088.88 800288 888008088808888088808080-2.z m 8803 8888 8.83.8 8.08.88 8.38.88 8.03.88 8088.88 820 888 888.0 8.088.88 8.088.88 - 8.080.88 820 08 8.8 8.888.88 8.088.88 8.088.88 88:08 88808800888808580-z.z 8.33 8888 8.8 88.8 8088.88 8.888 .88 8.3.8.88 8.38.88 88-88808588888388885 80803 8808 0.888.8 8.088.88 8.088.88 8.088.88 8.008.88 0:0 80803 8808 8.888.8 8.080.88 8.088.08 8.088.88 8.088.88 80208 88808080888800080-z.z .000 00 080580000 0805\80ox (88:000u08 0805\8000 008E< 808 800 808 808808 88 .000850 0000080000080nz.z 0880088008800 0500 08 00800008 80080008 00800080 800 0800080800 00800>800< .88 08888 107 0800000 08005080u000 800 00088008000 000000000 00000800000 088000 000 80800 000 080 .088 .000 0088 080 000 mm .*0 0800050800 0090 88.0- 88.0- 8.8 8.88 8.88 8.88 8.88 888080808888080-2-888088088-z 88.0- 88.0- 8.8 8.88 8.88 8.88 8.88 888088888880080-z-88xmao8880-2 88.0- 88.0- 8.8 8.88 8.88 8.88 8.88 888080808880080-z-880=880-z 80.0- 08.0- 8 8 0.8 8.0 8 8.08 8.0 8 0.88 8.0 8 8.88 8.0 8 8.08 888080888888080-z-88808-z 00m *0 .00 080588008 080888000 ~8u000u800< .00 08009 108 ground state are: (1) The degree of double-bond character in the C-N 223g. The double-bond character and the bond energy should be maxima when the oxygen, carbon, and nitrogen atoms of the amide framework are coplanar. Alternatively, if the carbon and nitrogen sigma bonds have spa hydridization the framework would be coplanar. This tends to stabi- lize the coplanar state. Inductive effects may alter the double-bond character by withdrawing or supplying electrons. Conjugation of the C-N bond with substituents will also have an effect. (2) Steric effects. Large, rigid substituents tend to force the trigonally hybridized atoms of the amide framework out of the plane. Repulsive forces increase as the substituent size increases, but may be relieved somewhat if the sub- stituents can be distorted or can rotate internally themselves to reduce their effective size. Westheimer and Mayer (119) have pointed out the importance of the bending or distortion of bonds in relieving steric strain. (3) Intermolecular interactions. A substituent group may inter- act with another molecule, for example, intermolecular hydrogen bonding. This could cause a variety of effects. The result might be to increase the effective size of the substituent group and thereby tend to force the trigonal atoms of the amide framework out of the plane. Alternatively, the association might cause the double-bond character to increase and thereby stabilize the planar ground state. The NMR spectra of N,N-dimethylformamide were analyzed by two dif- ferent methods: (1) complete neglect of the coupling between the N-methyl protons and the formyl proton; and (2) the introduction of the coupling constants into the theoretical lineshape equation. The results from the second method agree favorably with the results for N,N-dimethylformamide-dl in which the effect of this coupling is largely removed. The barrier in -.. :r 3...; ’ v -'I‘-. 109 N,N-dimethylformamide is considerably higher than in the remainder of the dimethylamides. The increased height of the barrier may be due to the effect of intermolecular hydrogen bonding which will stabilize the ground state (11,25,35) by formation of dimers or polymers. For N,N-dimethylacetamide, N,N-dimethylcarbamylchloride, N,N-di- methyltrichloroacetamide, and N,N-dimethylpropionamide (Table 37) the values of 46* agree with the best recent values within i 0.2 kcal/mole while E8 and AH; differ by as much as 1.0 kcal/mole, and AS* by as much as 4.6 eu, even though all were obtained by total lineshape analysis. These differences are not far from being within the sums of the random 40.1... errors and probably give a reasonable idea of the accuracy of the activa- tion parameters determined in this study. The effect of substituents appears to be, in nearly all cases studied, a lowering of the barrier to internal rotation. The effects are not large and separating out the steric, polar and resonance contributions is difficult (24). Both polar and resonance contributions should be small and similar for a series of saturated hydrocarbons substituents (120) and steric factors might therefore be expected to dominate. Larger substituents, whether on nitrogen or on carbonyl carbon, should tend to make the ground state more nonplanar and so reduce the barrier. In the solid state it has been shown that larger groups produce increasing non- ‘planar distortions in substituted amides by a twisting about the C-N loond (121). Solvent effects on AG:t have been shown to be small so one tnay be justified in comparing AG¢ for neat liquids (37). It has been pointed out (27) that AG¢ decreases with the size of the aliphatic group R in the series RCON(CH3)2. The new values for N,N-di- .Inethylacetamide and N,N-dimethylpropionamide fit neatly into this group, 110 for which AG* has the values 21.0 kcal/mole (R = H) (25), 18.1 kcal/mole (R 5 CH3), 17.2 kcal/mole (R = -C2H5), 16.2 kcal/mole (R = -CH(CH3)2) (122) and 12.2 kcal/mole (R = -C(CH3)3 (27). It was similarly shown in a study of the series CH3CONR'2 that the barrier decreases with increas- ing size of R' (28), with AB: = 17.7 kcal/mole (R' = - C2H5), 17.6 kcal/ mole (R' = isobutyl), and 16.2 kcal/mole (R' = isopr0pyl) reported. In the series of unsymmetrically N,N-disubstituted amides CH3CON(CH3)R' studied (Table 38) it is found that AG¢ decreases in the sequence 18.1 kcal/mole (R' = methyl), 18.0 kcal/mole (R' = ethyl), 17.9 kcal/mole (R' nrbutyl), 17.1 kcal/mole (R' = cyclohexyl), and 17.0 kcal/mole (R' of the steric and inductive substituent constants, E8 and 0* (120), are given in Table 38 and it is seen that AG$ decreases with increasing ability to withdraw electrons inductively and with increasing size (ex- cept for cyclohexyl). Both factors (more negative 0* and more negative E8 values) would be expected to lead to a decrease in AG:t for substitu- ents on nitrogen so the results are reasonable; however, one cannot judge their relative importance. Since values of ES may contain a hypercon- jugative component, and since the parameters are being used here in a somewhat different situation than that for which they were derived (120), any more detailed analysis of the correlations would not be justified. The values of K and of -AG for the conformational equilibria in 300° unsymmetrically N,N-disubstituted amides (Table 39) are somewhat lower than those obtained earlier from peak areas (49), but the errors in each case are rather large. Substituents decrease A6300, in the same order that they lead to decreases in the rotational barrier, but the energy changes are much larger for AG*. There must then be important substituent Fm“ ¢ isopropyl); AH and Ea decrease in the same order. Values 111 Table 39. Conformer equilibria in unsymmetrically N,N—disubstituted amides. Compound K=pA/pB A6300o K(peak areas)a (cal/mole) N-Ethyl-N-methylacetamide 1.04b -21 1.04 N-B-Butyl-N-methylace tamide 1.09 -50 1.13 rm" N-Cy clohexyl-N—me thy lacetami de 1.15 -81 1. 22 I N—IsoprOpyl-N-methylacetamide 1.25 -131 1.38 3The preferred configuration (A) has the larger alkyl group cis to carbonyl oxygen (49). ..uj The error in K is a few percent for each method. effects on the energy of the transition state. These would be expected to be quite different from those for the ground state since the configura— tion at nitrogen can be nonplanar as a result of the decreased delocali- zation of the lone pair. Resonance Assignments in Tertiary Amides by Lanthanide-induced Shifts The use of lanthanide complexes as reagents to induce paramagnetic shifts in the NMR spectra of polar organic molecules is now well estab- lished (84-101). The paramagnetic shift arises from the association of the lanthanide complex with the polar organic substrate and is largely dominated by dipolar (pseudo-contact) interactions (102,103). The mag- nitude of the paramagnetic shift is largely determined by the distance of the given nuclei from the lanthanide ion. Amides are known to protonate preferentially on the carbonyl oxygen (123), since the nitrogen lone pair 112 is extensively delocalized; one would, therefore, expect that the chemi- cal~shift reagent would be complexed through the lone pairs of electrons on oxygen and that the induced shifts in the gi§_(to carbonyl oxygen) and Egagg_N-alky1 substituents would differ. A linear correlation was found for the chemical shift for each group of protons in several amides and the mole ratio of shift reagent to amide. Chemical shifts relative to TMS were plotted against the mmle ratio of added shift reagent and values of the chemical shift for an equimolar ratio of shift reagent (5n=1 ) and pure amide (8n:O ) were obtained by CC14 CCl4 extrapolation. From these, values of AEu(AEu==5n=l ._8n=o ) were calcu- CCl4 CC14 lated for each group of protons (Table 40). Extrapolation of the lines to a molar ratio of zero allows assignments to be made for the amide resonances in the uncomplexed amide (Table 41). The validity of the method was established by the choice of l-methyl- 2-pyrrolidinone as a model system since the N-methyl group is fixed in the gig position. As expected, the gi§_methyl protons were shifted to a greater extent than the Egaggymethylene protons in the solutions con- taining the complexed amide. For the remaining amides, the resonances of the gi§f§£§£§_pair shifted the greater amount were assigned to the gig group. The assignments of Table 41 for the uncomplexed amides confirm those previously made by other methods for N,N-dimethylformamide (47,48,50,51, 124), N,N-dimethylacetamide (47,48,50,51,124), N,N-dimethyltrifluoroacet— amide (ll), N,N-diethylformamide (51), N,N-diisopropylformamide (51), and N-methyl-N-isopropylformamide (49). The method is rapid, readily appli- cable to a wide variety of amides and, to the extent of this investiga- tion, appears to be generally valid. When a single resonance is observed r11 113 Table 40. AEu values for several N,N-disdbstituted amides. 0 CH 10.12 0 CH 9.32 § / 3 \\ / 3 C‘-N C--N /’ ‘\t z’ \\ ‘\. //’ CH; 5.09 9.68 O CH CH 5.79 \ / 2 3 0 CH3 10.18 C--N §§ .z’ /’ ‘\. /’ \\ 11.39 H3O CH3 5.24 3.87 0 CH3 9.38 8.24 \ / /’ '\\ 8§§ x” F3C CH3 5.12 C-'N 1’ "x 10.81 H CH(CH3)2 2.14 0 CH 16.63 3.07 \ / 3 C--N ./’ \\ C13C CH3 7.99 0 CH3 9.81 §§bC Nil’ 0 CH3 11.65 / -‘ \ \C__N/ 13.82 H ca(c113)2 2.62 /’ \\ c1 CH3 6.18 4'14 \ / o mucus). 4.86 C-——N‘\‘ 8§§ ‘,/ CH3CH20/ CH3 6.24 /C—N\ CBoth isomers of CH0N(CH3)(i-03H7) are present in the same solution. Table 41. Proton resonance assignments in some N,N-disubstituted amides in CClu solution. 114 Compound N,N—Dimethylformamide N,N—Dimethylacetamidec N,N-Dimethyltrifluoroacetamided N,N-Dimethyltrichloroacetamide N,N-Dimethylcarbamylchloride N,N—Diethylformamide N,1‘\I-Diisopropylformamidee N-Methyl-N-isopropylformamidef Assignment of Upfield Resonance of cis-trans Paira 'Cflb cis cis cis cis trans 32 trans cis 32 -c_1_1_2 -c_1_1_ T°Cb 32 32 32 -22 trans 32 —cy_ -cu(cga)2 8The designation "cis" in this table indicates that the upfield resonance of the cis-trans pair for the indicated group is cis to oxygen in the CClu solution containing no shift reagent. bThe temperature was chosen, in each case, to be well below the coalescence temperature so that the isomers are distinguishable by NMR. CUse of Pr(fod)3 gave upfield induced shifts but the assignment of the pro- ton resonances was the same. din this amide the spin-spin couplings JH— JH_F(trans) (ll). eThis result is less certain than the others since the chemical shift between the methyl protons in the uncomplexed amide (2 1 Hz) is the same order of magnitude as the error in carrying out the extrapolation. f The major isomer (67%) has the N-methyl substituent cis to oxygen (49). are anomalous in that JH_F(cis) > 115 for the cis and trans protons this method can be used to determine whether the chemical shifts are accidentally equal or whether the amide is above the coalescence temperature. Carbon-13 Chemical Shifts of Amides The major advantage of carbon-13 over proton NMR for structure studies is in the increased separation of the chemically shifted peaks. Under normal operating conditions the linewidths in the spectra of the two nuclei are not very different whereas the shift range for carbon-13 nuclei is about twenty times larger than for protons. Since chemdcal shifts are sensitive to structural changes, the increased separation of, chemical shifts for carbon provides a more powerful way of differentiating proposed structures. The theory relating chemical shifts to structure is complex and, at present, does not allow quantitative interpretation of the experimental data. The most useful approach for structural deter- mination has proved to be a semi-empirical one based on the considerable body of experimental evidence, and supported in specific applications by the use of model compounds. The carbon-l3 chemical shifts in fifty mono- substituted and symmetrically and unsymmetrically disubstituted amides are given in Tables 32, 33, 34, and 36. The carbon-13 chemical shifts in amides are to a substantial degree influenced by inductive, resonance, and steric effects in a manner which is familiar for a variety of phenomena involving organic compounds.' The inductive effect of the nitrogen or carbonyl oxygen in amides on the chemical shifts of the carbons in substituents may be assumed to be inde- pendent of conformation but dependent on the degree to which each carbon 116 is substituted. The resonance effect relates to the kinds and stereo- chemical relationships between pairs of atoms on vicinal carbons and is assumed to affect only the shifts of the carbons directly involved. This effect, which has been considered by Grant t al. (57), is expected to be sensitive to the conformation. Steric effects have been shown to be very important in various kinds of hydrocarbons and are expected to be especially sensitive to conformational changes (57). The problem in sorting out these influences is that with many substitutions a composite of effects can be expected. Nevertheless, comparison of the chemical shifts in amides allows certain generalizations to be drawn. The carbon directly bonded to nitro- gen is shifted downfield by approximately 1.5 ppm whenever the formyl H proton is replaced by an alkyl group, probably as a result of an induc- tive effect which is transmitted by the electrons delocalized in the central C-N bond. A carbon directly bonded to nitrogen is also shifted downfield tn; about 9.0 ppm when the hydrogen bonded to nitrogen, as in a monosubstituted amide, is replaced by a carbon atom. This shift is almost identical to the negative inductive effect (+9.4 ppm) which is observed for hydrocarbons. The chemical shift difference between the carbons directly bonded to nitrogen which are gig and trans to the car- bonyl oxygen is consistently about 5 ppm for disubstituted formamides and decreases with the degree or kind of substitution on the carbonyl carbon. The large chemical shift difference arises from intramolecular electric and magnetic fields which are produced by the anisotropies of the C=0 and OC-R (R = H,CH3...) bonds (55). The chemical shift differences between cis and trans aliphatic carbons get smaller for carbons farther away from the carbonyl group. This is to be expected, since the anisotropy effect 117 falls off as the cube of the inverse of the distance from the center of thehanisotropic group. ,The monosubstituted amides exhibit only one set of carbon resonances for the N-alkyl substituent. Proton NMR studies,and a variety of other methods (114), have shown that the preferred configuration or isomer in monosubstituted amides is the one in which the N-alkyl substituent is cis to the carbonyl oxygen. In most of the disubstituted amides studied Ffimmfi the internal rotation about the central C-N bond can be slowed down on - l the NMR time scale by decreasing the temperature so that both rotational isomers could be observed. However, when the carbonyl carbon is bonded to a iybutyl group, decreasing the temperature to the freezing point _ "J of the particular pivalamide did not sufficiently slow down the rotation about the C-N bond so that both isomers could be distinguished. Assignment of the carbon-l3 chemical shifts for nitrogen substitu- ents were made with the aid of lanthanide shift reagents. The effect of the shift reagent in the carbon-13 spectrum of an amide was assumed to be the same as in the proton NMR spectrum, that is, the carbon is a gig: ‘ggggg pair shifted the greater amount can be assigned to the nitrogen substituent gig to the carbonyl oxygen. For the<1 and B carbons on nitrogen in N,N-di-gfpropylformamide the up-field carbon resonance for each gigfggggg pair was shifted the greater amount and was assigned to the carbon in the nitrogen substituent gig to the carbonyl oxygen. How- ever, the 7 carbons are reversed, the down-field carbon resonance for this gigfggggg pair being shifted the larger amount. § One of the purposes in obtaining the carbon-13 chemical shifts for a large number of amides was to use the data to derive a set of param- eters by which the chemical shifts in any amide could be predicted, 118 analogous to the method developed by Grant ggugi. (57) for hydrocarbons. Howgver, because of the inductive effects of both the nitrogen atom and the carbonyl group, and the large steric interactions between substitu- ents, a set of parameters could not be found with any reasonable certainty for the amides studied. SUMMARY Nuclear magnetic resonance methods have been employed in a series of F_—§ studies of the molecular structure and internal motions of amides in the liquid phase. (1) It has been shown that lanthanide shift reagents provide a re- liable method of assigning the NMR resonances to the gig_and giggg isos é;fij mers in disubstituted amides. Assignments have been made for a repre- sentative series of compounds. (2) The carbon-13 chemical shifts in fifty monosubstituted and di- substituted amides have been measured and empirically correlated with structural effects. (3) The barriers hindering internal rotation about the central C-N bond of a series of dimethylamides have been measured. The NMR method for carrying out these studies has been improved by the digitization of spectra followed by computer curve fitting of the lineshapes to the NMR parameters. (4) Internal rotation about the central C-N bond of a series of unsymmetrically N,N-disubstituted amides has been studied by the NMR technique developed in this thesis. The energies of activation have been related to properties of the substituents. 119 LIST OF REFERENCES I“? _a '2 _AIBCKE": r—‘w LIST OF REFERENCES 1. L. Pauling, "The Nature of the Chemical Bond," 3rd ed., Cornell University Press, Ithaca, N. Y., 1960, p. 281. 10. ll. 12. 13. 14. 15. 16. 17. O J. N. O J. J. C. Zabicky, Ed., "The Chemistry of Amides," Interscience, New York, Y., 1970. Ladell and B. Post, Acta. Cryst., l, 559 (1954). L. Katz, Dissertation Abstr.,‘i1, 2039 (1957). Pedone, E. Benedetti, A. Immirzi, and G. Allegra, J. Amer. Chem. Soc., 22, 3549 (1970). J. u. A. . C. Costain and J. M. Dowling, J. Chem. Phys., 3;, 158 (1960). H. Schwendeman, private communication. D. Phillips, J. Chem. Phys., g;, 1363 (1955). S. Gutowsky and C. H. Holm, J. Chem. Phys., 22, 1228 (1956). C. Woodbrey and M. T. Rogers, J. Amer. Chem. Soc., g3, 13 (1962). T. Rogers and J. C. Woodbrey, J. Phys. Chem., gg, 540 (1962). Allerhand, H. S. Gutowsky, J. Jonas, and R. A. Meinzer, J. Amer. Chem. Soc., gag, 3185 (1966). C. S. Johnson, Jr., in "Advances in Magnetic Resonance," Vol. I, J. S. Waugh, Ed., Academic Press, New York, N. Y., 1965, pp. 33-102. G. N. W. W. E. Stewart and T. H. Siddall, III, in "Progress in NMR Spectroscopy," Binsch in "Topics in Stereochemistry,” Vol. III, E. L. Eliel and L. Allinger, Eds., Interscience, New York, N. Y., 1968, pp. 97-192. E. Stewart and T. H. Siddall, III, Chem. Rev., 19, 517 (1970). Vol. VI, J. W. Emsley, J. Feeney, and L. H. Sutcliffe, Eds., Pergamon Press, London, 1969, pp. 33-147. H. Kessler, Agnew. Chem. Int. Ed., 9, 219 (1970). 120 _— 2" 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 121 J. P. Lowe in "Progress in Physical Organic Chemistry," Vol. VI, A. Streitweiser and R. W. Taft, Eds., Interscience, New York, N. Y., .1968, pp. 1-80. T. Drakenberg and S. Forsen, Chem. Commun., 1404 (1970). R. C. Neuman, Jr., and V. Jonas, J. Amer. Chem. Soc., 29, 1970 (1968). I“ W. Reeves, R. C. Shaddick, and K. N. Shaw, J. Phys. Chem., 12, 3372 (1971). L. W. Reeves and K. N. Shaw, Can. J. Chem., 39, 3683 (1971). L. W. Reeves and K. N. Shaw, Can. J. Chem., 32, 3671 (1971). ' A} R. F. Hobson, Ph.D. thesis, University of Waterloo, Waterloo, Canada, 1972. M. Rabinovitz and A. Pines, J. Amer. Chem. Soc., 2i, 1585 (1969). T. Drakenberg, K. Dahlqvist, and S. Forsen, Acta Chem. Scand., g3, 1 694 (1970). ' rggi L. L. Graham and R. E. Diel, J. Phys. Chem., 1;, 2696 (1969). T. H. Siddall, III, W. E. Stewart, and F. D. Knight, J. Phys. Chem., 13, 3580 (1970). H. S. Gutowsky, J. Jonas, and T. H. Siddall, III, J. Amer. Chem. Soc., 82, 4300 (1967). R. C. Neuman, Jr., and V. Jonas, J. Phys. Chem., 12, 3532 (1971). 0. A. Gansow, J. Killough, and A. R. Burke, J. Amer. Chem. Soc., 93, 4297 (1971). E. S. Gore, D. J. Blears, and S. S. Danyluk, Can. J. Chem., 33, 2135 (1965). A. G. Whittaker and S. Siegel, J. Chem. Phys., 3;, 3320 (1965). G. Fraenkel and C. Franconi, J. Amer. Chem. Soc., 8;, 4478 (1960). R. C. Neuman, Jr., and L. B. Young, J. Phys. Chem., 99, 2570 (1965). F. Conti and W. von Philipsborn, Helv. Chim. Acta., 29, 603 (1967). C. W. Fryer, F. Conti, and C. Franconi, Ric. Sci., 22 [2A], 788 (1965). D. G. Gehring, W. A. Mosher, and G. S. Reddy, J. Org. Chem., ii, 3436 (1966). A. Allerhand and H. S. Gutowsky, J. Chem. Phys., gi, 2115 (1964). 40. 41. 42. 43. 44. 45. 46. 47. d 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 122 R. C. Neuman, Jr., D. N. Roark, and V. Jonas, J. Amer. Chem. Soc., 22, 3412 (1967). 'c. Franconi, Scienza e Tecnica, N. s., 3, 183 (1960); 2. E1ektrochem., 92, 645 (1961). J. A. Weil, A. Blum, A. H. Heiss, and J. K. Kinnaird, J. Chem. Phys., ‘39, 3132 (1966). A. Mannschreck, Tetrahedron Lett., 1341 (1965). A. Mannschreck, H. A. Staab, and D. Wurmb-Gerlich, Tetrahedron Lett., 2003 (1963). If. .1 A. Calzolari, F. Conti, and C. Franconi, J. Chem. Soc. (B), 555 (1970). P. T. Narasimhan and M. T. Rogers, J. Phys. Chem., g;, 1388 (1959). v. J. Kowalewski and D. G. Kowalewski, J. Chem. Phys., _3_2_, 1272 j " (1960). L} J. V. Hatton and R. E. Richards, Mol. Phys., 2, 253 (1960); 5, 139 (1962). L. A. LaPlanche and M. T. Rogers, J. Amer. Chem. Soc., 22, 3728 (1963). F. A. L. Anet and A. J. R. Bourn, J. Amer. Chem. Soc., 21, 5251 (1965). M. Frucht, A. H. Lewin, and F. A. Bovey, Tetrahedron Lett., 1083, 3707 (1970). J. B. Stothers, Appl. Spectrosc., 22, l (1972). E. W. Randall, Chem. Brit., 1, 371 (1971). P. C. Lauterbur, J. Chem. Phys., 22, 217 (1957). W. McFarlane, Chem. Commun., 418 (1970). G. C. Levy and G. L. Nelson, J. Amer. Chem. Soc., 22, 4897 (1972). D. M. Grant and E. G. Paul, J. Amer. Chem. Soc., g9, 2984 (1964). T. D. Brown, Ph.D. thesis, University of Utah, Salt Lake City, Utah, 1965. J. D. Roberts, F. J. Weigert, J. I. Kroschwitz, and H. J. Reich, J. Amer. Chem. Soc., 22, 1338 (1970). R. Hagen and J. D. Roberts, J. Amer. Chem. Soc., 2i, 4504 (1969). 123 61. B. V. Cheney, Ph.D. thesis, University of Utah, Salt Lake City, Utah, 1966. 62. ”F. Bloch, w. w. Hansen, and M. Packard, Phys. Rev., Lg, 460 (1946). 63. H. S. Gutowsky, D. W. MCCall, and C. P. Slichter, J. Chem. Phys., 2i, 279 (1953). 64. H. M. McConnell, J. Chem. Phys., 22, 430 (1958). 65. K. J. Laidler, "Chemical Kinetics," McGraw-Hill, New York, N. Y., 1962. 66. S. W. Benson, "The Fundamentals of Chemical Kinetics," McGraw-Hill, { New York, N. Y., 1960. - 67. J. A. Pople, Disc. Faraday Soc., 22, 7 (1962). 68. J. A. Pople, W. G. Schneider, and H. J. Bernstein, "High-Resolution A ’7 Nuclear Magnetic Resonances," McGraw-Hill, New York, N. Y., 1959. ~ .5 69. A. Saika and C. P. Slichter, J. Chem. Phys., 22, 26 (1954). 70. M. Karplus and J. A. Pople, J. Chem. Phys., 22, 2803 (1963). 71. J. W. Emsley, J. Chem. Soc. (A), 1387 (1968). 72. J. E. Bloor and D. L. Green, J. Amer. Chem. Soc., 22, 6835 (1967). 73. H. S. Gutowsky and D. W. McCall, Phys. Rev., 22, 748 (1951). 74. H. S. Gutowsky, D. W. McCall, and C. P. Slichter, Phys. Rev., 22, 589 (1951). 75. E. L. Hahn and D. E. Maxwell, Phys. Rev., gg, 1246 (1951); 22, 1070 (1952). 76. N. F. Ramsey, Phys. Rev., 22, 303 (1953). 4 77. J. A. Pople and D. P. Santry, M01. Phys., 2, l (1964). 78. E. G. Paul and D. M. Grant, J. Amer. Chem. Soc., 22, 2977 (1964). 79. F. J. Weigert and J. D. Roberts, J. Amer. Chem. Soc., 22, 2967 (1967). 80. K. F. Kuhlmann, D. M. Grant, and R. K. Harris, J. Chem. Phys., 22, 3439 (1970). 81. D. Doddrell, V. Glushko, and A. Allerhand, J. Chem. Phys., 22, 3683 (1972). 82. K. F. Kuhlmann and D. M. Grant, J. Amer. Chem. Soc., 29, 755 (1968). 124 83. R. L. Lichter and J. D. Roberts, J. Amer. Chem. Soc., 22, 3200 (1971). 84.-C. C. Hinckley, J. Amer. Chem. Soc., 2i, 5160 (1969). 85. J. K. M. Sanders and D. H. Williams, Chem. Commun., 422 (1970). 86. J. Briggs, G. H. Frost, F. A. Hart, G. P. Moss, and M. L. Staniforth, Chem. Commun., 749 (1970). 87. G. H. Wahl, Jr. and M. R. Peterson, Jr., Chem. Commun., 1167 (1970). 88. J. Briggs, F. A. Hart, and G. P. Moss, Chem. Commun., 1506 (1970). 89. P. V. Demarco, T. K. Elzey, R. B. Lewis, and E. Wenkert, J. Amer. Chem. Soc., 22, 5734 (1970). 90. P. V. Demarco, T. K. Elzey, R. B. Lewis, and E. Wenkert, J. Amer. Chem. Soc., 22, 5737 (1970). 91. G. M. Whitesides and D. W. Lewis, J. Amer. Chem. Soc., 22, 6979 (1970). 92. c. c. Hinckley, J. Org. Chem., ;5_, 2834 (1970). 93. F. 1. Carroll and J. T. Blackwell, Tetrahedron Lett., 4173 (1970). 94. D. R. Crump, J. K. M. Sanders, and D. H. Williams, Tetrahedron Lett., 4419 (1970). 95. K. L. Liska, A. F. Fentiman, Jr., and R. L. Foltz, Tetrahedron Lett., 4650 (1970). 96. J. K. M. Sanders and D. H. Williams, J. Amer. Chem. Soc., 22, 641 (1971). 97. R. E. Rondeau and R. E. Sievers, J. Amer. Chem. Soc., 22, 1525 (1971). 98. W. Horrocks, Jr., and J. P. Sipe, III, J. Amer. Chem. Soc., 22, 6800 (1971). 99. J. Briggs, F. A. Hart, G. P. Moss, and E. W. Randall, Chem. Commun., 364 (1971). 100. E. Wenkert, D. W. Cochran, E. D. Hagaman, R. B. Lewis, and F. M. Schell, J. Amer. Chem. Soc., 22, 6271 (1971). 101. M. Witanowski, L. Stefaniak, H. Januszewski, and Z. W. Wolkowski, Tetrahedron Lett., 1653 (1971). 102. R. J. Kurland and B. R. McGarvey, J. Magn. Res., 2, 286 (1970). 103. B. R. McGarvey, J. Chem. Phys., 22, 86 (1970). 104. R. F. Nystrom, W. H. Yanko, and W. G. Brown, J. Amer. Chem. Soc., 29, 441 (1948). 3"”? -—- "(A- .A. 125 105. J. D. Cox and R. J. Warne, Nature, 165, 553 (1950). 106. 6. M. McElvain and C. L. Stevens, J. Amer. Chem. Soc., 22, 2667 (1947). 107. L. A. LaPlanche, Ph.D. thesis, Michigan State University, East Lansing, Michigan, 1963. 108. J. C. Woodbrey, Ph.D. thesis, Michigan State University, East Lansing, Michigan, 1960. 109. W. C-T. Tung, Ph.D. thesis, Michigan State University, East Lansing, Michigan, 1968. 1*‘1 110. D. DeMaw, Ed., "The Radio Amateur's Handbook," 47th Ed., 1970, p. 244. 111. L. N. Mulay in "Treatise on Analytical Chemistry," Pt. I, Vol. IV, E I. M. Kolthoff and P. J. Elving, Eds., Interscience, New York, N. Y., ,. 1963, p. 1827. _J 112. J. L. Dye and V. A. Nicely, J. Chem. Ed., 22, 443 (1971). 5:; 113. T. Drakenberg, K-I. Dahlqvist, and S. Forsen, J. Phys. Chem., 12, 2178 (1972). 114. L. A. LaPlanche and M. T. Rogers, J. Amer. Chem. Soc., 22, 337 (1964). 115. S. Glasstone, K. J. Laidler, and H. Eyring, "The Theory of Rate Processes," McGraw-Hill, New York, N. Y., 1941, p. 296. 116. V. N. Kondrat'ev, "Kinetics of Chemical Gas Reactions," Book 1, Academy of Sciences of U.S.S.R., Moscow, 1958. Translated by U.S. Joint Publications Research Service, New York, N. Y., 1962. 117. R. J. Kurland and E. B. Wilson, Jr., J. Chem. Phys., 22, 585 (1957). 118. R. J. Kurland, J. Chem. Phys., 22, 2202 (1955). 119. F. H. Westheimer and J. E. Mayer, J. Chem. Phys., 13, 733 (1946). 120. J. Shorter, Quart. Rev. (London), 22, 433 (1970). 121. P. Ganis, G. Avitabile, S. Migdal, and M. Goodman, J. Amer. Chem. Soc., 22, 3328 (1971). 122. G. Isaksson and J. Sandstrom, Acta Chem. Scand., 2i, 1605 (1967). 123. R. J. Gillespie and T. Birchall, Can. J. Chem., 3;, 148 (1963). 124. B. B. Wayland, R. S. Drago, and H. S. Henneike, J. Amer. Chem. Soc., 22, 2455 (1966). 1111111111le 8 2 2 2 3 8 O 3 o 3 S" H” II " fl '1'. ”WIN/,1