CARBONUEE MAGNENC RESQNANCE OF SOME ALWHAHC AME PHENYL CQMPOUNDS AND $QW‘ENT swag-Es Q‘F fiakfi‘i ETHYL CGMPQUMDS USING E’ROTQN MAGNET“: RESQNAMCE Thesis for Hm Degree 0? pk. D. MICHIGAN STATE UNIVERSITY Roy Lee Fciey 1964- _. “IBRARY ‘ ‘ Michigan 5m H University L — .3 ‘ £13.... MIC!‘IIQTT‘~I S'TATE L'L’IVERSITY DEPARTMENT OF: CI-IEIVIISTRY EAST LANSING, MICHIGAN A BST RACT CARBON- 13 MAGNETIC RESONANCE OF SOME ALIPHATIC AND PHENYL COMPOUNDS AND SOLVENT STUDIES OF SOME ETHYL COMPOUNDS USING PROTON MAGNETIC RESONANCE by Roy Lee Foley The carbon-l3 magnetic resonance spectra of several aliphatic, phenyl, diphenyl, and triphenyl compounds were obtained using a Varian high-resolution nuclear magnetic resonance spectrometer operating at a fixed radiofrequency of 15. 085 Mcps. The carbon-13 spectrawere obtained for all compounds in natural abundance using 15 mm. external diameter sample tubes. The low natural abundance of the carbon-13 isotOpe and its longer relaxation times necessitated the use of rapid passage dispersion mode conditions to obtain spectra. The solvent studies on the ethyl compound were performed by using a Varian A'-60 spectrometer to obtain the proton spectra. The carbon-13 chemical shifts were found to vary over a wide range in contrast to fluorine and proton chemical shifts. The carbon- 13 chemical shifts are interpreted in terms of inductive and anisotropic effects and are shown in many cases to correlate quite well with atomic or group electronegativities. In addition, the correlation of carbon- 13 chemical shifts with molar susceptibility data, supports the contention of Pople that the paramagnetic contribution to the chemical shift is the dominant contribution for carbon-13 shifts. The carbon-l3 proton and carbon-13 fluorine spin- spin coupling constants have been recorded for all compounds studied. In the o. -substituted toluenes very good Roy Lee Foley agreement between the spin-spin coupling constants calculated using zeta values and measured coupling constants was observed. A study of proton n.m. r. spectra of a series of ethyl compounds M(CZH5)n in various solvents indicates that inter- and intramolecular interactions for solute and solvent are absent. CARBON-13 MAGNETIC RESONANCE OF SOME ALIPHATIC AND PHENYL COMPOUNDS AND SOLVENT STUDIES OF SOME ETHYL COMPOUNDS USING PROTON MAGNETIC RESONANCE BY Roy Lee Foley A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1964 ACKNOWLEDGMENTS The writer wishes to express his sincere appreciation to Professor Max T. Rogers for the supervision, encourage- ment, and suggestions given during the formation of this thesis. He also wishes to express his gratitude to the National Science Foundation and the Atomic Energy Com- mission for grants subsidizing this research. >:< * >;< >1< >:< >:< >:< >{z >:< >:< >:< >',< >;< >:< >:< ):< ii TABLE OF CONTENTS INTRODUCTION . . . . . . . . . HISTORICAL REVIEW 0 O O 0 O O O O O O 0 THEORY O O O O O O O O O O O Lamb's Theory of Chemical Shifts Ramsey' 3 Theory of Shielding Constants . An Approximate Theory of Chemical Shifts. . A General Molecular Orbital Theory of Chemical Approximation A Approximation B Approximation C Approximation D Approximation E Approximation F Atomic Contributions to Chemical Shifts ' Molecules............. Carbon-Carbon Single Bonds . Carbon-Carbon Double Bonds. Carbon-Carbon Triple Bonds . Allenes........... . Ca rbonyl Groups . Nuclear Spin-Spin Interaction . . Nuclear Relaxation . . . . . . . . . LineShapes............. Line Broadening. . . . . . . . . . . Line Asymmetry. . . . Saturation . . . . . . . Double Irradiation. . . EXPERIMENTAL . . . . . . Spectrometers. . . . . Compounds Studied . . 0 iii 0 O O O Shifts . Page 16 16 16 17 18 18 21 22 23 26 28 29 31 33 33 34 TABLE OF CONTENTS - Continued Determination of Spectral Parameters . . . . . . . . . . . 38 Proton Magnetic Resonance. . . . . . . . . . . . . . 38 Carbon-13 Magnetic Resonance. . . . . . . . . . . . 38 RESULTS....... ...... ....... 41 Effects of Solvent on the N.M. R. Spectra of Some Organo- metallic Compounds . . . . -. . . . . . . . . . . . . 41 Substituted Acetic Acids . . . . . . . . . . . . . . . . . . . 44 Some Fluorinated Aliphatic Compounds .......... 55 a-Substituted Toluenes . . . . . . . . . . . . . . ..... 66 Diphenyl Compounds . . . . . . . . . . . . . . . . . . . . 86 Triphenyl Compounds. . . . . . . . . . . . . . . . . . . . 93 A1cohols..... ...... ................109 DISCUSSION OF RESULTS. . . . ................. 122 Solvent Studies of Ethyl Compounds ...... . . . . . 122 Correlation of Chemical Shifts with Molar Susceptibilities 123 Correlation of Chemical Shifts with Electronegativity . . . 123 SUMMARY. 0 O O O O O O O ...... O I O 0 O 0 0 0 O O O O O O 126 BIBLIOGRAPHY. . . . . . ....... . . . . . . . . ..... 127 iv LIST OF TABLES TABLE Page I. Physical Constants of Alipahtic Compounds Studied. . . 35 II. Physical Constants of u-Substituted Toluenes Studied. . 36 III. Physical Constants of Diphenyl Compounds Studied . . . 36 IV. Physical Constants of Triphenyl Compounds Studied . . 37 V. Solvent Effects on the Carbon-13 Chemical Shifts of CarbonDisulfide..................... 40 VI. Proton Chemical Shifts in Ethyl Compounds . . . . . . 41 VII. Spin-Spin Coupling Constants of Pb(CHZCH3)4 in CC14 . . 42 VIII. Spin-Spin Coupling Constants of Sn(CHZCH3)4 . . . . . . 42 IX. 15.1 Mc./sec. Carbon—13 Spectra of Some Substituted AceticAcids........................ 54 X. 15.1 Mc./sec. Carbon-13 Spectra of Some Fluorinated AliphaticCompounds................... 69 XI. Carbon-13 Chemical Shifts and Spin-Spin Coupling Con- stants of Some Aliphatic and Fluorinated Aliphatic Compounds................‘........ 71 XII. 15.1 Mc./sec. Carbon-13 Spectra of Some Substituted TOIueneSO . O O O O O O O O O O O 0 O O O O O O O O 0 O O O 90 XIII. 15.1 Mc./sec. Carbon-13 Spectra of Some Diphenyl Compounds........................ 97 XIV. Chemical Shifts (ppm.) and Spin-Spin Coupling Con- stants (cps.) for Some Diphenyl Compounds. . . . . . . 98 LIST OF TABLES - Continued TABLE XV. XVI. XVII. XVIII. Page 15.1 Mc./sec. Carbon-13 Spectra of Some Triphenyl Compounds. . .. . . . . . . .. . . . . . . . 107 Chemical Shifts (ppm.) and Spin-Spin Coupling Con- stants (cps.) of Some Triphenyl Compounds ...... 108 Chemical Shifts for Some Aliphatic Alcohols . . . . . . 110 Calculated and Observed Spin-Spin Coupling Constants in Substituted Toluenes. . . . . . . . . . . . . . . . . 111 V1 LIST OF FIGURES FIGURE . The shapes of n.m. r. signals: (a) Slow passage absorp- tion, (b) Slow passage dispersion, (c) Rapid passage dispersion......... ..... 15-1Mc./sec. carbon-13 spectra of benzene . . . . . Internal chemical shifts versus mole fraction of solute forethylcompounds. . . . . . . . . . . . . .. . . . . 15.1 Mc./sec. carbon-13 spectra of trifluoroacetic aCid. O O O O O O O O O O O I O O O O O O O O O O O O 0 15.1 Mc. /sec. carbon-13 spectra of difluorochloro- acetic a'CidO 0 O O 0 O O 0 O 000000 O O I O O O O O 15.1 Mc./sec. carbon-13 spectra of fluorodichlor'o- acetic acid OOOOOOOOOOOO O O O O O O O O O 0 15.1 Mc./sec. carbon-13 spectra of trichloroacetic aCi-d. O O O O O I O O O O 0 O O O O O O O O O O O O O O O O 15.1 Mc. /sec. carbon-13 spectrum of dichloroacetic aCidO O O O O O O O O O O O O O O O O G O O O O 0 O O O 15.1 Mc. /sec. carbon-13 spectrum of dichloroacetic aCido O O O O O O O O O O O O O O O O 0 O O O O O O O O 15.1 Mc./sec. carbon-13 spectra of chloroacetic acid . 15. 1 MC. /sec. carbon-13 spectrum of phenylacetic aCid. O I O O O O O 0 O O O O O O 0 O O O O O I 0 O O O 15. 1 Mc./sec. carbon~l3 spectrum of phenylacetic aCidO O O O O O O 0 O O O O O O O 0 O O O O 0 O O O 0 0 vii Page 27 3O 43 45 46 47 49 50 50 51 52 53 LIST OF FIGURES - Continued FIGURE 11(a). 11(b). 12(a). 12(b). 13(a). 13(b). 14. 15(a). 15(b). 16(a). 16(1)). 17(a). 17(b). 18(a). 18(b). 15.1Mc./sec. carbon-13 anhydride. . . . . . . . . 15.1Mc./sec. carbon-l3 anhydride. . . . . . . . . 15.1Mc./sec. carbon-13 15.1Mc./sec. carbon-13 15.1Mc./sec. carbon-13 acetylacetone. . . . . . . 15.1Mc./sec. carbon-13 acetylacetone. . . . . . 15.1Mc./sec. carbon-13 l, 1, 2, 2-tetraf1uoroethane . 15.1Mc./sec. carbon—13 ethanol.......... 15.1Mc./sec. carbon-13 ethan01.. . .. . .. . . 15.1Mc./,sec. carbon-13 2-propanol . . . . . . . 15.1Mc./sec. carbon-13 Z-pl‘OpanOI o o o o o o o o 15.1Mc./sec. carbon-13 2-methy1-2-propanol . . 15.1Mc./sec. carbon-13 2-methyl-2-propanol. . . 15.1Mc./sec. carbon-13 15.1Mc./sec. carbon-l3 spectrum of t rifluoroac etic spectrum of trifluoroacetic spectrum of acetylacetone. . spectrum of acetylacetone. spectrum of hexafluoro- spectrum of hexafluoro- spectra of 1, 2, -dibromo- spectrum of 2, 2, 2-trifluoro- spectrum of 2, 2, 2-trif1uoro- spectrum of 1,1, l-trifluoro- Spectrum of 1, 1, 1-trif1uoro- O O O O 0 O O O O 0 O O O 0 O spectrum of 1, 1, latrifluoro- spectrum of 1, 1, 1-trif1uoro- O O O O O O 0 ° 0 0 0 O O O 0 O spectrum of toluene. . . . . spectrum of toluene. . . . . viii Pace 56 56 57 58 59 6O 62 63 64 65 65 67 68 73 74 LIST OF FIGURES —' Continued FTGURE 19(a). 19Ib). 20(a). 20 (b) . 21(a). 21(b). 22(a). 22(b). 23(a). 23(b). '24(a). 24(b) . 25. 26. 27. 28. 29. 30. 31. 32. 15. .1 1 15.1 15.1 15.1 15.1 15.1 .1 15.1 .1 h4c./sec. L4c./sec. LAc./sec. hAc./sec. bAc./sec. hAc./sec. L4c./sec. h4c./sec. LAc./sec. /sec. h4c./sec. LAC. hAc./sec. h4c./sec. hAc./sec. L4c./sec. L4c./sec. bAc./sec. bAc./sec. /sec. LAc./sec. hdc. carbon—13 carbon—13 carbon-13 carbon-13 carbonu13 carbon~13 carbon-13 carbon-13 carbon-13 carbon—13 carbon—13 carbon-13 carbon-13 carbon-13 carbon-13 carbon-13 carbon-13 carbon-13 carbon~13 carbon-13 spectruni spectrunn spectruni spectrun1 spectrun1 spectrunn spectruni spectrunn spectrun1 spectrunn spectrun1 spectruni Page of benzylbromide. . of benzylbromide. . of benzylchloride. . Cd benzylchloride. . ofbenzyhnercapuni ofbenzyhnercapuni of diphenylmethane. of diphenylmethane. of benzylcyanide . . of benzylcyanide . . of benzylamine. . . of benzylamine . . . spectra of benzylalcohol . . . spectra spectra spectra spectra spectra spectra spectra ix of diphenylether . . . of diphenylsulfide . . of diphenylselenide. . ofchphenyhnercury of triphenylamine . . of triphenylarsine . . 75 76 77 78 80 81 82 83 84 85 87 88 89 92 94 95 96 100 of triphenylphosphine 101 102 LIST OF FIGURES - Continued FIGURE 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 15.1 Mc./sec. carbon-13 spectra of triphenylstibine. . 15.1 Mc./sec. carbon- 13 spectra of triphenylbis- mUthine. . O O O O O O I O O O O O I O O O O O O O O O O O 15. 1 Mc. /sec. carbon- 13 spectra of triphenylmethane . Chemical shifts versus molar susceptibilities for some aliphaticalcohols..................... CX3 chemical shifts versus molar susceptibilities in some substituted acetic acids . . . . . . . . . . . . . . Chemical shifts of carbonyl carbon atoms versus molar susceptibilities for some substituted acetic acids. . . . The aliphatic carbon atom chemical shifts versus molar susceptibilities insome phenylmethanes. . . . . Chemical shifts of ortho carbon atoms versus molar susceptibilities for some diphenyl and triphenyl com- pounds. 0 O I O I O O O I O O. O O O O O O O O O I O O O O CX3 chemical shift versus group electronegativity for a series of substituted acetic acids. . . . . . . . . . . Chemical shifts of carbonyl carbon atoms versus group electronegativity for a series of substituted acetic aCids. O O O O O O O O O O O O O O O I I O O O O O O O O O CHZX chemical shift versus group electronegativity in substitutedtoluenes. .. . .. . . ...... .... . Chemical shifts of ortho carbon atoms versus Pauling's electronegativity of central atoms for some diphenyl and triphenyl compounds. . . . . . . . . . . . . . . . . Page 104 105 106 113 114 115 116 117 118 119 120 121 INTRODUCTION The applications of nuclear magnetic resonance spectroscopy to the study of molecular structure are now well established. However, it has only recently become possible to successfully apply this technique to the carbon- 13 nuclei in natural abundance. The long thermal relaxation times (T1) and low natural abundance of the carbon- 13 isotope have tended to discourage the initiation of a compre- hensive study of carbon-13 resonances in its compounds. The one per cent natural abundance of the carbon-13 isotope is just sufficient to allow its effects to be detectable in the proton and fluorine spectra of simple molecules and, more important to permit the carbon-13 spectra themselves to be observed. The measurement of peak positions and spin-spin coupling constants may provide a useful aid in the analysis and understanding of molecular structure. The research reported in this thesis consists of the following: (1) an examination of the carbon-13 magnetic resonance spectra of several aryl derivatives in order to ascertain which features of the observed spectra may be correlated with known atomic and molecular parameters, (2) an examination of the carbon-13 magnetic resonance spectra of several fluorinated aliphatic compounds in order to determine the effect of progressively replacing protons by fluorine atoms, and (3) an examination of the proton magnetic resonance spectra of some ethyl derivatives in carbon tetrachloride, cyclohexane, and chloroform solution to determine possible effects of solvent on the internal chemi- cal shifts and spin- spin coupling constants. HISTORICAL REVIEW Most of the nuclear magnetic resonance (n. m. r.) data reported in the literature to date have been concerned with nuclei whose natural abundances are approximately one hundred per cent, mainly hydrogen, fluorine, phosphorus, and boron. The commercial avail- ability of sensitive high- resolution instruments has now made it possible to observe carbon—13 satellites in hydrogen (1, 2) and fluorine (3,4) spectra. The first observation of carbon-13 spectra in natural abundance was reported by Lauterbur (5). Employing a transmitter frequency of 8. 5 Mc/sec. and a stationary magnetic field of 7940 gauss, Lauterbur was able to obtain the carbon-13 chemical shifts and hydrogen-carbon-13 spin- spin coupling constants of several simple aliphatic compounds. From this study Lauterbur concluded that the spin- spin coupling between carbon- 13 and directly bonded hydrogen is comparable in magnitude to that between hydrogen and other nuclei, and tends to in-- crease with decreasing magnetic shielding of the proton. Holm also reported some of the earlier carbon-13 chemical shifts and spin-spin coupling constants (6). The results of Holm essentially substantiated the conclusions of Lauterbur. Some comprehensive natural abundance studies of carbon- 13 shielding in aromatic systems have been made. In a study of aromatic hydrocarbons (7) it has been found that the variations in pi-electron densities are primarily responsible for the variations in the chemical shifts in aromatic systems. Lauterbur (8) has concluded from a study of phenols, anisole and dimethoxybenzenes that the methyl effects on carbon-13 chemical shifts and coupling constants are relatively small. 2 The correlation of carbon-13 chemical shifts with electronegativity for substituted aliphatic and aromatic compounds have been well illus- trated by Lauterbur (9). The analysis of carbon-13 spectra has been accomplished in three ways. Lauterbur (7e12,) has applied the method of methyl substi- tution to the analysis of spectra of benzene derivatives. Spiesecke and Schneider (13, 14) have taken advantage of the fact that when deuterium is substituted on a carbon in the benzene ring no peak will appear for that carbon. Double irradiation methods have been applied by Friedel and Retcofsky (15) in analyzing the carbon-13 spectra of olefins and other hydrocarbons. Karplus and POple (16) have recently published a theoretical treat- ment of carbon-13 n.m. r. chemical shifts in conjugated molecules. The empirical relation ACT-A = QIPA'I) suggested by Lauterbur (7), and by Spiesecke and Schneider (14), has been derived using molecular orbital theory. AU'A is the chemical shift. with reference to benzene, (1 is a positive constant that is approximately 160 ppm. , and P is the pi electron density. In addition, the shielding A is shown to be a function of the free valence of the atom under consider- ation and of the polarity of the sigma bonds. Savitsky and Namikawa (87) have shown that carbon- 13 chemical shifts in CH3X, CZH5X, i- C3H7X, and t-C4H9X compounds are additive. Their results indicate that the carbon-13 chemical shift toward lower field per methyl substitution depends very little on the nature of X for fi-carbons but vary considerably for a-carbons. Additivity relations of the carbon-13 chemical shifts in para-disubstituted benzenes have also been found by Savitsky (88). These relations are helpful in the assignment of the carbon-13 n.m. r. spectra of polysubstituted ben- zenes. Savitsky and Namikawa (89) have used carbon-13 n.m. r. to investigate geometric isomerism about the ethylenic double bond. It has been concluded that the carbon- 13 chemical shift differential between cis and trans carbons could not be satisfactorily explained solely in terms of long range anisotropy effects. An alternative explanation involving possible deviation from coplanarity of bulky cis substituents was pre sented. THEORY Lamb's Theory of Chemical Shifts Chemical shifts have their origin in the magnetic screening of the nucleus which arises from the orbital electronic currents induced by an external magnetic field. These currents also give rise to diamagnetic polarization (17). When the external magnetic field is H the total magnetic moment of the induced current is XmHo where o’ Xm is the molecular diamagnetic susceptibility. The secondary magnetic field due to the induced currents at any given nucleus is -()‘iHo where -0‘i is the magnetic screening constant. The total field experienced by a given nucleus, which determines the n.m. r. frequency, is given by H. = HON-041). (1) 1 The chemical shift may then be defined as (18) 5=_i____1: (2) where Hi is the resonant field of the signal being measured at a fixed frequency and H1. is the corresponding field for a given reference. 7 From equation (1) it is seen that the differences in the shieldings of various nuclei are reflected by the differences in the screening constants. There have been several attempts to calculate the screen- ing constants of atoms and molecules. Using a classical approach, Lamb was able to derive an expression for the screening constants in atoms (19). The resultant expression is generally known as Lamb's formula 07- - 313—.— 2 (-.~) (3) 3m-1c Ti where (-e) is the charge on the electron, mi is the electronic mass, c is the velocity of light, and ri is distance of the ith electron from the nucleus, the sum is over all the electrons. Equation (3) gives the diamagnetic contribution of the electrons to the magnetic field and is applicable only to atoms since it depends on the spherical symmetry of the electric field of the nuclear electric potential. Ramsey's Theory of Shielding Constants Ramsey has made a quantum mechanical calculation of molecular screening constants (20). Using second-order perturbation theory Ramsey derived an equation for the molecular screening constants as 0'= .)(0|2;3.r|o>-3AE(012—41——km) (4) 3mc‘z k where m0 j is the total angular momentum and AB is the average value of all the electronic exitation energies. The first term of equation «4) corresponds to Lamb's expression for the diamagnetic shielding of single atoms. The second term arises from the lack of spherical symmetry of the electric potential and is called paramagnetic term. The most serious limitation of Ramsey's equation is that for large molecules the diamagnetic and paramagnetic terms turn out to be comparable in magnitude and opposite in sign thereby tending to cancel each other. An Approximate Theory of Chemical Shifts The theory of chemical shifts has also been developed in terms of breakdown of diamagnetic currents into a sum of atomic terms as was first proposed by Saika and Slichter (21). The total screening constant for any atom A may be broken down into the following four terms. AA AA AB A ' G’A:0’d +g’p +2: 0'“ +0” ’rmg (5) B(#A) , AA . . . The first term, O'd , gives the contr1bution to the secondary magnetic field at nucleus A due to the diamagnetic currents on atom A itself. For an isolated spherical atom, it is the only contribution and is given explicitly by the Lamb's formula of equation (4). The second term, prA is the contribution due to the para- magnetic currents on atom A which gives the susceptibility term. This term was first calculated for fluorine atoms by Saika and Slichter (21) who showed that it was much more sensitive to chemical structure than the first term. The total variation of O-pAA in going from F2 to a highly ionic system is of the order of 10-3. For most other atoms variations in this term will give the main contributions to the chemical shifts. The hydrogen atom is the principle exception since the absence of low lying p—orbitals makes this paramagnetic term negligibly small. The third term, TAB, accounts for the contribution to the screening of atom A by the atomic circulations on atom B. When the magnetic effects of these neighboring currents are treated in a dipole approximation, this term will include only the local anisotropy of the local susceptibility on atom B (22, 23). If atom A is on or near the axis of high diamagnetism of B, there will be an increase in the average screening. This effect falls off as the inverse cube of the A-B inter- nuclear distance and is not likely to exceed 10"5 in magnitude (l7). A . The fourth and final term of equation (5), 0‘ ’ r1ng, takes into consideration the contribution to the screening due to ring currents which cannot be localized on any atoms. The magnitude of this term is usually less than 2 x 10'6 but it plays an important role in determin- ing the spectra of aromatic compounds (24, 25). A General Molecular Orbital Theory of Chemical Shifts An approximate theory of molecular diamagnetic susceptibilities based on the use of gauge-invariant atomic orbitals (GIAO) has been recently proposed by Pople (28). The theory is developed in terms of an independent electron model and is applicable to the theory of chemi- cal shifts. The motion of each electron will be given by a one-electron Hamiltonian A )2 + V (6) where A is the magnetic vector potential, P is the momentum, and V an appropriately averaged effective electrostatic potential. When a molecule is placed in an external magnetic field _I-_1_ with a single magnetic nucleus that has a magnetic moment at the origin, A will be given by é=iEXE+exz/r3- (7) With a magnetic field present, the molecular orbital may be written as a linear combination of atomic orbitals Vi = ‘3 C’iu X 11 (8) where X11 is a gauge-dependent atomic orbital (17) that may be repre- sented as Xu=¢uexp(-%Efi '3) (9) ¢u is the normal atomic orbital and A is the value of the vector potential at the nuclear position of ¢u . The use of the variation principle (26) will yield linear'equations of the form . = . 1 EHquiv £1 E SUV Civ ( 0) where 6i is the eigenvalue of the secular equation and represents the energy possessed by an electron in the molecular orbital. The matrix elements Huv and Suv are the coulombic, resonance, and overlap integrals of molecular orbital theory (27) and are given by H 11V I 713mg? I exp { ($114.. zévl’f- I ‘1’: 5%(3‘15‘5'5) }z+g¢v¢r c (11) II and :9: SUN: f x“ Xydxr -.-. f exp { (TIE-”Au .. fiVI' I} ¢u*¢vd,r (12) It may be seen from equations (11) and (12) that the Hamiltonian and overlap integrals depend on the magnetic field because of the gauge factors. The following approximations are those suggested by Pople in the development of the theory of molecular diamagnetic susceptibilities (28). Approximation A The exponential part of equations (12) and (13) .is taken outside of the integral sign and I. is replaced by )- (Ru-I» Ev): where Eu and Ev 10 are the nuclear positions of the atomic orbitals ¢u and ev, respectively. Replacing r by%- (Ru-t RN) is required only when the atomic orbitals are not on the same atom and does appear to be reasonable since the midpoint between nuclei u and v will be approximately the center of the "overlap distribution. " This approximation does give rise to the possibility. that the matrix Huv may be non-Hermitian. Should this possibility occur it is to be resolved by splitting the resulting matrix into Hermitian and non-Hermitian parts with retention of only the Hermitian pa rt. . e Equat1ng Luv = (mHAu - év). (Ru + RV) (13) the matrix elements Huv and Suv become _ , * 1 e 2 2H...- exp “Luv’ 14>. [‘75) I '3 + 216-" (1.21 +311 I’vd'r , * l e z + exp (IL...) I .v [(75) {1: + flé'éd) + Xl¢ud7 (14) and SW = exp (111W) f q): ¢v d7. (15) Approximation B The assumption is made that the magnetic field is weak enough for all magnetic energies to be small compared with the energy differences between molecular orbitals Ei - Ej' This approximation allows the magnetic effects to be treated as perturbations and all matrix elements can be expanded in orders of magnitude in A _ (0) (1) (Z) , , , H - H... + H... + H... + “6' s = 5(0) + 5(1) + s”) uv uv uv 11V IIV + o . o (17) 11 more specifically I . 0 Hss= + -——t 1.1 Herm (Z) (0) ' * H... =% Li... H..* ($32) Luvtf (hue-Aw: ¢vdrl (20> He rm 2 e * z +2mc2{f¢u (xi-6'11) ¢VdT1Herm where H (0) is the Hamiltonian in the absence of a magnetic field. For the overlap matrix elements (0) __ * Suv ‘ I 4’u ¢v d7“ (21) Sm = 1L 5(0) + 5“” (22) uv 11V uv uv 5“” = - 5- 1.2 5(0) + 1L 5‘01 +, 5(0) (23) 11V 1.1V 11V uv 11V uv Approximation C The atomic orbitals on the same atoms are assumed to be mutually orthogonal and the overlap integral between atomic orbitals on different atoms is negligible. The result of this approximation is to make the overlap matrix take the form S = 6 (24) and make it independent of the magnetic field. The latter part of this approximation, while not quantitatively accurate, leads to a major simplification that is usually made in molecular orbital theories of non- magnetic phenomena . 12 The following two approximations are made to simplify the matrix elements Hiilv) and HLZV) (2). Approximation D The integral f (I): (é-éu)’ I: ¢vdT is to be neglected if dpu and rbv are not on the same atom. This approximation may be justified as follows: In a uniform magnetic field, the operator (é-év)' P is pro- portional to the orbital angular momentum operator at the center of the nucleus v. When ¢v is an atomic s orbital the integral vanishes. When ¢v is a p or d orbital the operator converts it into another p or d orbital. This results in an overlap-type integral between different atoms which, in accordance with approximation C, is negligible. The (Z) to result of this approximation is to cause the second term in Huv vanish as Luv vanishes, if the integral does not. Approximation E >:< The integral f ¢u (A-Av)z¢vdr is negligible when ¢u and ¢v are not identical. In addition to the five preceding approximations for the theory of molecular diamagnetic susceptibility, it is convenient to make an additional approximation for the theory of chemical shift. Apm'oximation F The local vector potential (é'év) is replaced by the local vector potential of a magnetic field that is given by -1 é'S—v_TI-_—va(£'B-v) (25) (‘1 (‘1 (‘1. 13 In view of the foregoing approximations, it is seen that the break- down of the molecular screening constants into intra-atomic terms is accomplished by neglecting certain integrals involving atomic orbitals on more than one atom. The diamagnetic contribution to the suscepti- bility, X3, arises from the gauge factor modification of the atomic orbital, equation (9). The paramagnetic contribution to the susceptibility comes from changes in the LCAO coefficients which corresponds to the mixing of the ground and certain excited states of the molecule by the magnetic field. The explicit relations for 7C: and X: in the z direction are as follows (17): A e‘2 A (7‘ d)zz = - W i puu( x2 + 3’2qu (26) (A: 1 — 4:122:2TAE QA (27) QA = Z QAB (28) CAB: PYAPxBaaAB-PYAPYB) + PYAPYB (zaAB- PXAPXB) + ZPXAXBPYAYB (29) where the various Puv elements are components of the charge-density and bond-order matrix for the unperturbed molecule, OCC P = 2 z c. c (30) uv Iu iv and the sum is taken over occupied molecular orbitals. The para- magnetic part, equation (27) has been. simplified by assuming that the excitation energies required in the perturbation theory can be replaced by an average value AE. The result can then be expressed in terms of 14 the Puv matrix for the molecular ground state; PxA PXB is the element of the matrix for the 2px atomic orbitals on atoms A and B. When A = B,“ PXA PXB the bond order between the two atomic orbitals. is the charge density in ZPxA orbital and when A 71 B, The explicit relations for the components of the magnetic screen- ing constant are: (1) Local Diamagnetic Currents -1 o-rd = ‘3——2‘ E Puu (r )uu (31) This is similar to the Lamb formula derived from classical con- sideration and is easily calculated if the LCAO molecular orbitals are known. (2) Local Paramagnetic Currents AA 0.) : — ZN"l < r'3 > (32) p 2p 7‘19 Here < r‘3 >2p is the mean value of r'3 for the 2p atomic orbital. The direct proportionality to the local paramagnetic contribution 7( should lead to a correlation of the chemical shift with diamagnetic susceptibility data for atoms with 2p electrons. The < r'3 > proportion- ality factor makes the local paramagnetic contribution large and is likely to be dominant in the application to carbon, nitrogen, oxygen, and fluorine. (3) Currents on Neighboring Atoms AB_1 _ B 5 0" ~19: N 1:126 1a,, (3RBQRBB - R2B6I1IB)/ RB (33) 15 Here a and 13 are the tensor suffixes and R o‘(a =x, y, z) are the B components of the vector from nucleus A to nucleus B. If the B B susceptibility on atom A is isotropic (X 05 = x 6 QB) this term is zero so the contribution is frequently called the neighbor anisotropy effect. When the local susceptibility tensor on atom B is axially symmetric O’AB takes the form J’AB=%-(A7LB)(1-3cosz{B)/R3B (34) where A XB(= XE - XiB ) is the anisotropy of the susceptibility and YB is the angle between the axis of anistropy and the AB internuclear line. A theory of carbon—13 n.m. r. chemical shifts in conjugated molecules has been developed recently by Karplus and Pople (16). The theory shows that the paramagnetic contribution to the carbon-13 chemi- cal shifts is the dominant one. The derived expressions for the para- magnetic contributions to the shielding in conjugated molecules are AA 1132 4 p = — [fizz-rIAEII 2p [2 + I; (~1—- FA” (35) AA__[eZBZ (AE3)] (1:3) [z+—4i)t(1-PAA)+-é(~f—-F) {p _ ch 2P 9 H z z 9 A (36) where FA if the free valence of the atom and AH is the polarity para- meter. Equation (35) defines the paramagnetic contribution for the carbon atom which is in a planar conjugated system with sp‘2 hybridization and bonded to three other carbon atoms. Equation (36) gives the para- magnetic contribution for the carbon atom which is in a planar conjugated system with sp2 hybridization but bonded to two carbon atoms and one hydrogen atom. 16 Atomic Contributions to Chemical Shifts in Organic Molecules Pople's theory can be used to make approximate estimates of atomic contributions to the diamagnetic susceptibility for some of the more important organic groups. For atoms other than hydrogen, the paramagnetic contribution X: is the most sensitive to electronic structure and estimates of the relative shifts of carbon and the neighbor anisotropy effect for hydrogen can be made by considering this term only. Carbon-Carbon Single Bonds The theory predicts that all carbon atoms in saturated compounds have the same X: (6.46 x 10'6) and are isotropic. Therefore all carbons in unstrained paraffins should have the same chemical shift. The experimental chemical shifts for paraffinic carbons are spread over a range of approximately 30 p.p. m. to low field from methane. Although these shifts have not been satisfactorily explained they may be due to some delocalization of the bonding electrons. Since the theory predicts that the carbon atoms are isotropic, calculated chemical shifts for all paraffinic hydrogens are the same. There is a range of approximately 1. 5 p.p.m. to low field from methane in the experimental data, o’(-CH) < M-CH.) <0’ (-CH.) < men.) but this may be due to changing electron density. Ca rbon- Ca rbon Double Bond 8 When the magnetic field is perpendicular to the carbon-carbon double bond, the theory predicts larger 7C: terms on the doubly-bonded (‘3 7" H4 (7 In 17 carbon atoms for ethylene and other olefins. This contribution arises from the mixing of the 04—? 1t and 17 +fexcited states with the ground state by the magnetic field. The calculated and observed chemical shifts for ethylene relative to ethane are given in Table I (17). Table 1. Calculated and Observed Chemical Shifts For Ethylene 6 6 Atom (Calculated) (Observed) (p-p-m-I (p-p.m.) Carbon -46 - 120 Hydrogen -2.1 -4. 2 The calculated chemical shifts are in qualitative agreement with the observed low diamagnetism of olefins and the low field shift of carbon resonance from ethane. The calculated shifts are in the right direction but are somewhat smaller than the observed values. Carbon- Ca rbon T riple Bonds The theory predicts that the acetylene molecule will, have a high diamagnetic anisotropy along the molecular axis. This should cause a high-field shift for the proton due to the neighbor anisotropy effect. The fact that the acetylene proton resonance occurs to the low field of ethane (-0. 6 p. p. m.) can be interpreted as a cancellation of neighbor anisotropy effect by the low-field shift due to reduced charge density on the hydrogen atom. The carbon chemical shift in acetylene should be close to that of the paraffins. The observed carbon shifts lie at an intermediate position between singly and doubly-bonded carbons. 18 Allenes Calculations on the allene molecule are of particular interest since they give an additional paramagnetic term A X: which is twice as large on the central atom as on the two end atoms. This is due to the fact the additional term in the plane of the double bond occurs for both directions perpendicular to the C=C=C axis for the central atom, but only for one direction for the end atoms. Hence it is to be expected that the carbon resonance of the central atom should be shifted to low field relative to the end atoms. The experimentally observed carbon shifts for allene do exhibit the separations of the kind predicted. However the observed difference in chemical shift between middle and end carbon atoms in allenes is 140 ppm. whereas the calculated separation is 40 ppm. The calculated shifts again are too small. Carbonyl Groups The theory provides an interpretation of the low diamagnetism of aldehydes and ketones and also predicts anisotropies for the carbon A and oxygen atoms. Table II summarizes the paramagnetic terms X p (17). Table II. Calculated Paramagnetic Contributions to the Atomic Molar Susceptibilities for the Carbonyl Group (Units of 10) (X) (X) (X) Y’l: , pXX p yy p ZZ X c 6.46 10.76 6.46 0 10.83 10.00 4.31 /C=O On the carbon atom the largest paramagnetic contribution is in the y- direction and the anisotropy is comparable to that for carbon atoms in 19 X P)zz for oxygen can be interpreted similarly. The high paramagnetic term the carbon-carbon double bond. The difference (Xp)yy - ( for oxygen in the x- direction arises from the low lying n -—> 11* excited state, which is mixed with the ground state by a magnetic field in this direction. From Table II it can be seen that the anisotropies for both atoms lead to negative (low field) contributions to the screening con- stants of aldehydic protons. The total calculated neighbor anisotropy effect is -2.6 p. p.m. of which -2. 2 p.p.m. comes from carbon. The observed chemical shift of aldehydic protons relative to paraffinic hydrogen is -8 p. p.m. It may be that the theory underestimates the anisotropies in the carbonyl group. The main advantages of the foregoing theoretical developments lie in the fact that all the contributions to the screening are deduced within a single theoretical framework and that the theory is particularly suitable for the chemical shifts involving atoms of hydrogen, carbon, nitrogen, oxygen, and fluorine. Equation (32) should be a suitable starting point for discussing correlations between chemicals shifts and atomic contributions to diamagnetic susceptibility. Hameka, who has developed an exact theory of the diamagnetic susceptibilities for diatomic molecules from SCF-LCAO-MO functions by introducing GIAO (29, 30), has raised some serious objections to Pople's approximate theory (31). Hameka contends that the deviations caused by approximations A-D are not negligible but that they are small when compared with the errors introduced by approximation E. In fact, Hameka claims that the approximation E is incorrect and concludes that it is generally not possible to express a molecular diamagnetic susceptibility as a sum of atomic contributions since interatomic terms are of major importance (31). 20 Pople has answered Hameka's objections by pointing out that part of the error caused by neglect of the two-center integral has been compensated for by associating larger coefficients with the one- center terms (32). In addition, Pople has emphasized that the theory based on atomic orbitals has a decided advantage in the application of n. m. r. chemical shifts, since the magnetic field is measured at the center of a particular atom. This approximate theory has been successfully applied by Pople to the hydrogen and carbon chemical shifts in carbon-carbon single, double, and triple bonds, allenes, and carbonyl groups (17). In addition, Karplus and Pople (33) have used a similar approach to develop a theory of carbon-13 n.m. r. chemical shifts in conjugated molecules. The theory does provide an interpretation of some of the empirical features of carbon-13 chemical shifts (33) in spite of the severity of the approximation. It should be emphasized that Pople's theory is to be taken as a step to simplify another complex problem and is of consider- able value in this respect. The theory of chemical shifts has also been developed from the valence bond approach. Some recent contributions to the theory of shielding constants include those of O'Reilly (34) who used a group theoretical method to calculate the chemical shifts of nuclei of ions in crystalline fields. Kern and Lipscomb (35, 36) have calculated the shielding constants in diatomic molecules by simplifying the paramagnetic term of Ramsey's equation with a judicious choice of the gauge of the vector potential and an approximation involving cancellation of average excitation energies of the matrix elements of the linear and angular momentum operators. In addition Hameka (37), and Ghosh and Sinka (38), have surveyed the various valence bond approaches used in calcu- lating shielding constants. Shielding constants in aromatic systems 21 have been investigated by Maddox and McWeeny (39), Hall and Hardesson (40), and by Jonathan, Gordon and Dailey (41). Nuclear Spin—Spin Interaction In 1951 Gutowsky, McCall and Slichter observed that high resolu- tion spectra frequently exhibited hyperfine structure which, in contrast to the linear dependence of chemical shifts, was independent of the applied magnetic field (42, 43, 44). Similar effects had been observed by Hahn and Maxwell in the modulation of the spin-echo envelope in pulse experiments (45, 46). It was concluded that the observed effects were caused by an indirect coupling of the nuclear momentsii, which is transmitted from nucleus to nucleus by the paired electrons comprising the valence bonds. The spin- spin coupling constant between two nuclei of spins i and j has the form (47): Jij = X, If Ki, (37) Where Ji’ is the spin-spin coupling constant, T is the magnetogyric ratio,‘his Plancks constant divided by 2 17 , and K is a constant depend- ing upon the molecular electronic structure. The magnitude of the coupling constants for hydrogen are small (0-20 cps.) but may persist over several chemical bonds (48, 49). The coupling constants between other nuclei increase with atomic number and may be as large as several kilocycles (50). Recently the problem of the sign of the spin-spin coupling constants has received considerable attention. Using double irradiation or spectrum analysis, the relative but not the absolute, signs of spin-spin coupling constants may be obtained. A knowledge of the absolute sign of spin- spin coupling constants is of importance for comparison of theory with experiment. Karplus (51) has suggested one method for the obtainment of 22 absolute signs which is based on the fact that the theory of spin- spin coupling for directly bonded atoms is on much more solid ground than that for more widely separated atoms and predicts a positive sign for directly-bonded atoms. Karplus has further suggested that in organic compounds the carbon-13-proton coupling provides a suitable reference for determining absolute signs of proton-proton coupling constants. Such absolute signs determinations have been carried out by Lauterbur and Kurland (52) and Anet (53) who found only partial agreement with the expectations from valence bond theory. The current theoretical treatment of spin- spin coupling in organic molecules suggests that there exists a correlation between the gem proton coupling constants and the H-C-H angle and between vicinal coupling constants and the H-C-C-H dihedral angle. A number of authors have measured such spin- spin coupling constants in a number of relatively simple molecules of known conformation, and compared the results with theoretical predictions. These include Graham and Rogers (54), Gutowsky and Juan (55), Williamson and Johnson (56), Anet (57), Barfield and Grant (58), and Bernstein and Sheppard (59). Only partial agree- ment between theory and experiment is generally found, and the actual spin- spin coupling constants are no doubt modified by substituent effects and other factors (60). Nuclear Relaxation When a collection of nuclei is placed in a magnetic field there will be an equilibrium distribution of these nuclei into the various spin states. For simplicity, nuclei with only two spin states will be con- sidered. The populations of the two spin states will not be equal, but will follow the Boltzmann distribution law with the state of lower energy being more populated (61). With the absorption of energy from the 1_, 0* .h- .c 7'" ' A“! Ti; 23 magnetic field, nuclei in the lower states will go into the upper state. There will also be a tendency for nuclei in the upper state to return to the lower state. The latter process is of extreme importance in the theory of nuclear magnetic resonance and is usually referred to as the relaxation process. The two general types of relaxation are the spin- lattice and the spin— spin relaxation processes. The spin~lattice or longitudinal relaxation, (T1), involves the establishment of thermal equilibrium between a collection of nuclear magnets with different quantum numbers and results in the establish- ment of an equilibrium value of the nuclear magnetization along the magnetic field axis (62). Among the interactions which contribute to T1 are (a) dipole-dipole interactions, (b) nuclear dipole-other magnetic moments interations such as electrons and molecular magnetic moments, and (c) nuclear quadrupole-electric field gradients. The spin-spin or transverse relaxation (T2) represents the loss of phase coherence of the precessing collection of magnetic nuclei. The interactions contributing to T2 are (a) magnetic field inhomogeneity, (b) direct dipole-dipole coupling in solids, and (c) spin- spin interactions. Line Shapes The quantitative relationships between the shape of an n. m. r. signal and the relaxation times T1 and T; can be obtained by solving the Bloch equations (63) ' _ . MX Mx - Y (MYHO + MLH1 Slnwt) — 7f;— (38) M=f(MHcoswt-MH)-Mz- (39) y Z I. X 0 T2 . _ M - M2 = X ("MxHi sin wt - MyI-Ii cos wt) ___er_‘;1\_4_0__ (40) 24 where Ma( 0. = x, y, z) is the component of the magnetic moment per unit volume, K is the magnetogyric ratio, 1:10 is the applied magnetic field, Ii; is a field perpendicular to I_-I_o rotating with angular frequency 00, and M0 is the equilibrium value of Mz. The Bloch equations can be simplified by referring them to a set of axes rotating with the applied field I_-1_1 rather than to the fixed axes x, y, and z. This transformation will give a new set of axes rotating with angular velocity 6) about the axes. The new components of M which are parallel (in phase) and perpendicular (out of phase) to the direction of H1, are called (.1 and v components. These components have the following relations. Mx =11 COS wt - V Sinwt (41) My: -usin wt-vcoswt (42) or alternately, C.“ II Mx cos cot - My sin wt (43) v=-stinwt-Mycos wt (44) Substitution of equations (41) and (42) into the equations (38), (39) and (40) gives the following form for the Bloch equations: 11 +-*-+va=0 (45) T2 . v v+—-Awu+wl=0 (46) T2 Mz+h—4-zf'fl9— -o.)v=0 (47) 1 where 25 Au) : COO-(.01, XHO '3 (.00, and r141: (.01. The steady state solutions of equations (45), (46), and (47) can be obtained by setting all time derivatives equal to zero. The steady state solutions are 2 L01 T2 A0) = M u 0 1+T§ A00 + wlleTz (48) V - MO (.01 T2 (49) 1+T§Aw + wfirlrz Under experimental slow passage conditions, {1 , v, M; fr; 0, the rate of absorption of energy is proportion to v and is given by equation (49). If the magnitude of the oscillating field is small such that a? T11“, << 1 (50) then the rate of absorption of energy becomes proportional to 2T2 1+ TE Am: (51) which gives the line-shape function 2T 2 (52) 3M = 1+417sz(}’0- )f)r Equation (52) describes the shape of the signal for the absorption or v mode, Figure 1a. Under appropriate experimental conditions it is possible to observe u, the in phase component of the magnetization. Under the conditions of slow passage the line shape function is proportional to equation (49). 26 If saturation is negligible, that is, when equation (50) holds, the dis- persion mode line-shape function is If...)r g(v) z ”4.213 (r.- r)? (53) which is shown in Figure lb. Figure 1c illustrates the line shape for the rapid passage dis- persion mode. Pure rapid passage conditions are not always achieved experimentally and the dotted line in Figure 1c illustrates the line shape under conditions intermediate between rapid passage and slow passage in the dispersion mode . Line Broadening The absorption of a n.m. r. signal occurs over a range of fre- quencies rather than a single frequency thus causing the signal to be broadened. There are a variety of factors contributing to the broaden- ing of a signal. Among these is the natural line width due to spontaneous emission which is determined by the finite lifetime in the upper state. The line widths are affected by the spin—lattice relaxation process. The order of magnitude of the broadening can be approximated by the uncertainty principle. AEAt .1: H (54) where AE and At is the uncertainty in energy and time, respectively. From equation (52) it can be seen that the uncertainty in frequency of absorption is (217At)"l and the line width measured in terms of frequency is proportional to the reciprocal of T1. In solids and highly viscous liquids, where the nuclei remain in the same relative positions for a long period of time, the interactions 27 Figure 1. The shapes of n.m.r. signals: (a) slow passage absorption, (b) slow passage dispersion, (c) rapid passage dispersion. 28 between magnetic dipoles can lead to a greater broadening than that caused by T1. The broadening is caused by the spin— spin relaxation time, T2, and is of the order of 105 cycles/sec. which is substantially larger than normal values of l/Tl. In liquids and gases the molecules are rotating rapidly and the dipole-dipole interaction is effectively averaged out. Under these circumstances T1 and T3 are approximately equal in magnitude. When nuclei have spin greater than one-half their electric quadropole moments will interact with the electric field gradients. This will give rise to additional mechanisms by which relaxation will occur. The net effect will be to decrease the value of T1 and T2 causing additional broadening. The essential instrumental limitation which contributes to the broadening of n.m. r. signals is the inhomogeneity of the static magnetic field. Ho, over the dimensions of the sample. This effect is of major importance in carbon- 13 natural abundance spectra where it is customary to use non- spinning sample tubes of large external diameter. Of particular importance in carbon-l3 spectra studies is the broadening caused by unresolved spin- spin couplings with distant hydro- gen atoms. These can be resolved in a few carbon- 13 spectra and have been measured in proton (65, 66) and fluorine spectra (67, 68); variations of these distant couplings will result in changes in the widths and heights of peaks and will affect intensity measurements which are usually based solely on peak heights. Line Asymmetry In addition to those factors contributing to the shapes and broaden- ing of n.m. r. signals discussed in the previous sections, there are factors which are of particular significance to natural abundance carbon-13 studies. These factors affect the symmetry of the carbon-L13 resonance signal and are discussed below. 29 There is an inherent asymmetry in the rapid passage dispersion mode carbon-13 spectra. In Figure 2 are shown the 15.1 Mc./sec. spectra of benzene under such conditions. The inversion of the spectrum with reversal of sweep directions is to be noted. It has often been 0 bserved that the appearance of the carbon-l3 spectra exhibits a strong dependence on the sweep direction. This makes it desirable that all spectra be presented for both sweep directions. Another source of asymmetry of carbon-13 spectra is the magneti- zation-transfer effect. This causes the second peak in the respective sweep directions to be decreased in intensity by an amount which depends upon the sweep rate and the relaxation times of both carbon-l3 and hydrogen. Similar effects are observed for higher multiplets. The magnetization-transfer effect is caused by a rapid transfer of non- equilibrium nuclear spin magnetization from one magnetic environment to another (69). Magnetization transfer in chemical reactions has been studied in detail by McConnell and Thompson (70). However, transfer by relaxation of spin-coupled nuclei although quite similar has not been treated quantiatively to date (71). Saturation The absorption of energy from the radio frequency field will not only cause a reduction in the population of nuclei in the lower state but also a reduction in the probability for further absorption. The magnitude of the latter will increase as the amplitude of the oscillating field in- creases. This effect which is known as saturation can be defined in terms of a saturations factor, Z, given by 1 z = (1 + VHETJZ)’ (55) 3O Ho Increase \ HO decrease T 7 Figure 2. 15.1 Mc./sec. carbon-13 spectra of benzene. 31 It is to be emphasized that in equation (55) the greater the radio frequency power applied to the sample, the greater the degree of saturation. In addition, the combination of a large radio frequency field, long spin- lattice relaxation times, and narrow absorption lines may prevent the observation of a given line or spectrum. This is of particular significance in natural abundance carbon-13 studies where the long relaxation times of the carbon-13 nuclei and the large radio frequency power customarily used necessitate special precautions. Double Ir radiation The hyperfine splitting due to nuclear spin-spin interactions between nuclei possessing magnetic moments can be removed by double resonance. This can be accomplished by simultaneously irradiating one set of coupled nuclei with a strong radio frequency field close to its resonance frequency while observing the second set of nuclei with a weak radio frequency field. The essential objective is to cause the nuclei irradiated by the strong radio frequency field to undergo rapid transitions which reduces the lifetimes in states with a given m value. When the lifetime becomes much less than J‘l, the group observed with the weak field is effectively decoupled. The double resonance technique is more easily applied to nuclei with different magnetogyric ratios than when nuclei have the same magneto- gyric ratio. Hence, carbon-13-proton decoupling is more readily obtained than proton-proton decoupling. The theory of double irradiation has been developed on a rigorous statistical quantum mechanical basis by Bloch (72). An approximate theoretical treatment, limited to the case where J << 6 has been given by Arnold (73). More recently the theory of double irradiation has been treated by Freeman and Whiffen (74), Anderson and Baldeschwieler (75), and Anderson and Freeman (76, 77). The latter papers are far more 32 comprehensive and contain graphs and tables from which frequencies and intensities in double irradiations spectra for Aan(m, n f 3) can be obtained. EXPERIMENTAL Spectrometer s The Varian A-60 Spectrometer was used to obtain the proton magnetic spectra. The internal chemical shift measurements for the ethyl compounds were obtained using 250 or 100 cps. sweep widths. The probe temperature was approximately 350C. The carbon-13 magnetic resonance spectra were obtained using a Varian Associates high-resolution nuclear magnetic resonance spectrometer, model V-4300-2. A Varian Associates model V-4331A RF probe was used to obtain spectra at a fixed frequency of 15. 085 Mc. /sec. The constant magnetic field was obtained with a Varian Associates model V-4012A twelve-inch high-resolution electro-magnet equipped with a model VK-3513 field trimmer and a model V-4365 field homo- geneity control unit; the power supply for the electromagnets was a model V—21OOA power supply equipped with a field reversing mechanism. The stability of the spectrometer system was enhanced by use of a Varian Associates model VK-3506 magnetic flux stabilizer. Further stability was achieved by using a Sorenson 10005 a. c. voltage regulator to regulate the line voltage to the spectrometer console. The magnetic cooling system, consisting of distilled water circulated through the cooling coils and recycled through a refrigerated copper tank, was maintained at a temperature of 19. 7 i 0. 30C. The room temperature was regulated by a commercial air—conditioning unit. Calibration of the spectra was obtained by using audio frequencies from a Hewlett-Packard Model 200 CD wide range oscillator, to modulate the field. Modulation of the field by audio frequencies produces side-bands 33 34 on either side of the main signal. The separation of the side—bands from the main peak is equal to the modulation frequency. A Hewlett- Packard Model 521~A electronic counter was used to measure the audio modulation frequencies. The spectra were recorded on a Varian Associates Model G-lO Graphic recorder. The probe insert had an internal diameter of 16 mm. which permitted the use of 15mm. external diameter sample tubes. All spectra were obtained under rapid passage dispersion mode conditions. The sweep rate employed was approximately 300 cps. per minute. Since the line widths increase with increasing radio frequency power H1, a compromise had to be reached between the signal strength and resolution in a given spectrum. It has been empirically established that better spectra are usually obtained when the radio frequency field is at a setting of coarse-2 and fine-5. To obtain the dispersion mode spectra, the probe was tuned at maximum radio frequency power. The R—F field was then set at coarse-2 and fine-5. After minimizing the direct current signal level meter with the fine red paddle (u-mode) the fine blue paddle (v-mode) was turned counter clockwise to introduce a leakage of 5 to 10 microamperes. The resultant direct current signal level meter reading was usually 50 to 60 microamperes. Compound 5 Studied All compounds studied in this research are listed in Tables I, II, III, and IV. The compounds were purchased from Eastman Organic Chemicals, Peninsular ChemResearch, Inc. , Columbia Organic Chemi- cals Co. , Inc. , and The Matheson Company. The solvents carbon disulfide and chloroform were J. T. Baker AR grade and were used with- out further purification. The carbon- 13 spectra of both solvents exhibited no impurity peaks. Table I lists the aliphatic compounds studied. 35 The compounds were dried for at least twenty-four hours over Drierite and distilled at atmosphere pressure. Chloroacetic acid and tri- chloroacetic acid which are solid at room temperature were distilled without drying. Table 11 lists some physical constants of toluene and its derivatives. Tables III and IV list some physical constants of the diphenyl and triphenyl compounds, respectively. Table I. Physical Constants of Aliphatic Compounds Studied Compound Boiling Point (OC.) Molecular Weight oncogn 71. 2 114 CFZC1COZH 119.7 130 CFClzCOzH 160. 3 147 CC1,cozH 195.4 163 CHClzCOZH 193. 8 129 CHZC1COZH 187. 9 95 ' (CF3CO)ZO ‘ 38. 2 210 (CF3CO)ZCHZ 63. 7 208 (CH3CO)2CHZ 174. 6 100 .CFzBrCFzBr 45. 3 260 CF3CHZOH 72. 5 100 CF3CH(OH)CH3 76. 7 114 CF3C(CH3)ZOH 80. p—I 128 (I- If}. 36 Table II. Physical Constants of a—Substituted Toluenes Studied Compound Boiling Point (0C) Molecular Weight ¢CH3 110. 6 92 ¢CH2C1 176. 3 127 ¢CHZBr 196. 8 171 ¢CH2NH2 182. 4 107 ¢CHZCN 230.6 117 ¢CHZSH 193.5 124 ¢CHZOH 203. 2 108 ¢CHZ¢ 263. 1 168 Table III. Physical Constants of Diphenyl Compounds Studied Compound Boiling Point (OC.) Molecular Weight (1)20 255. 7 170 (1’25 293. 9 186 ¢zSe 297. 4 233 ¢zHg 12381 355 aMelting point. 37 Table IV. Physical Constants of Triphenyl Compounds Studied Compound Melting Point (°c.) Molecular Weight ¢3N 125. 5 245 ¢3P 79. 2 262 63As 60. 7 306 4135b 51.9 353 e381 76.5 440 ¢,CH 92. 8 244 Sample Preparation The samples for which proton spectra were to be obtained were degassed and sealed in 5mm. external diameter A- 60 tubes. For the oxygen sensitive ethyl compounds the solvent was added to the sample tube and the solute was then vacuum distilled into the sample tube. For the carbon- 13 studies the liquid samples were mixed with CS; to give thirty to fifty volume per cent solutions which were placed in a 15mm. external diameter sample tube filled to a height of 20 to 30 mm. , and the tubes sealed under vacuum. Alternately, the neat liquids were placed in a 15mm. external diameter sample tube attached to a 14/35 standard-taper female joint. A 14/35 standard-taper male joint sealed at one end was used as a stopper. The internal diameter of the male joint was reduced to 5 mm. This permitted a 5mm. sample tube containing 57. 3% carbon— 13 enriched methanol to be placed within the sample tube to be used as a secondary standard. Where solubility was sufficient to give a good carbon-13 signal, solids were dissolved in carbon disulfide to yield a saturated solution. The saturated solution 38 was then sealed under vacuum in the 15 mm. sample tube. Compounds having low solubility in carbon disulfide were dissolved in chloroform and the spectra obtained using the above standard-taper tubes. Determination of Spectral Parameters Proton Magnetic Re sonanc e To obtain the internal chemical shifts and spin- spin coupling con- stants for A383 and A3B2X systems the experimental spectra must be compared with the theoretical spectra. This is necessary because the ethyl compounds considered in the solvent studies do not have first order spectra at 60. 000 mcs. The theoretical spectra were obtained from the table of line frequencies and relative intensities for the A3Bz system listed by Corio (48), with one modification. Corio gives the data for the cases where the internal chemical shifts are positive--that is, when the resonance of the methyl protons, A group, occur at a higher field than the resonance of the methylene protons, B group. For the systems considered the resonance peak for the methylene protons occurs at a higher frequency than that of the methyl protons, resulting in a negative internal chemical shift. This requires a change of sign for the frequency of each line position. The internal chemical shifts were obtained by measuring the frequency separation between the A6 line and the mean of the B4 and B5 lines. The values for the spin- spin coupling constants were obtained by simultaneous solution of two corres- ponding equations for the line frequencies. Carbon-l3 Magnetic Resonance The inversion of rapid passage dispersion mode spectra with re- versal of sweep direction and the strong dependence of the appearance of a spectrum upon the sweep direction have made it necessary to record 39 the spectra in both sweep directions. The direction of increasing field will have the peaks pointing up and the higher applied fields to the right. The direction of decreasmg field has the peaks pointing down and the higher applied fields to the left. For both spectra, the direction of sweep is from left to right, increasing in the first case and decreasing in the second. The measurements of peak positions were accomplished by measuring equivalent points in spectra taken with increasing and decreasing sweep field and averaging the results. The spectra were calibrated using a 1000 cps. modulation frequency. A secondary cali- bration used was the spacing between the two outer members of the quartet from the enriched sample of methanol. All chemical shift measurements are given using carbon disulfide as the internal and external standard. This follows the convention established by Lauterbur who has shown that solvent effects on the chemical shifts of carbon disulfide are negligible. Table V gives the solvent effects on the chemical shifts of carbon disulfide in various solvents (78). Due to the negligible solvent effects on the chemical shift of carbon disulfide no corrections for the diamagnetic susceptibility have been made. The assignments of the peaks in the aliphatic compounds can be made on the basis of the multiplicity caused by the spin- spin interactions. This was possible because the overlapping of peaks occurred infrequently. One exception was (CF3CO)ZCHZ where the spin-spin coupling caused over- lapping peaks. The assignment of peaks was greatly aided by the unambiguous assignments that had been made for (CH3CO)ZCHZ. For the a- substituted toluenes the aromatic portion of the spectrum consists of two or three peaks. The five ring carbon atoms can be assigned by consideration of the magnetization transfer effects and relative peak heights. Table V. Solvent Effects on the Carbon-13 Chemical Shifts of Carbon Disulfide'" 6 Solvent (ppm.) Corrected (ppm.) Nitrobenzene O. 0. 5 Acetone 0. 0-0 Aniline O. 0. 2 Cyclohexane 0. 0.1 Carbon Disulfide (0 (0) Bromoform 0. O. 4 3): Reference (78). Note: All solutions contained one-third carbon disulfide by volume. The maximum deviation of an individual measurement from the average was 0. 2. The assignments of peaks occurring in the spectra of the diphenyl and triphenyl derivatives are somewhat more difficult to make. The height of each peak was measured in both sweep directions and averaged. The assignments for diphenyl ether were based on the similarity of the carbon- 13 spectra of phenol which was analyzed by Lauterbur with the aid of methyl substitution (8). For the remaining phenyl compounds the assignments were made on the basis of the factors discussed previously. In addition, quadrupole broadening and larger spin values for the arsenic, antimony, and bismuth (I: g- ; %,%~; 521-) compounds will cause the substituted carbon peaks not to be observed. The para carbon atoms can usually be distinguished from the ortho and meta carbon atoms by relative peaks heights . could not be made with complete certainty. The choice between ortho and meta carbons However, Spiesecke and Schneider have shown on the basis of deuterium substitution that the range of chemical shifts for meta carbons is usually very small, 2.4 ppm. RESULTS Effects of Solvent on the N.M. R. Spectra of Some Organometallic Compounds The results from the solvent studies of the n.m. r. spectra of a series of ethyl compounds are listed in Tables I and II, and Figure 1. Table VI gives the internal chemical shifts of the pure compounds and also their internal chemical shifts at infinite dilution. Figure 1 is a graph of chemical shift against mole fraction of solute from which the internal chemical shifts at infinite dilution were ob- tained. The variation of spin-spin coupling constants with concen- tration for all of these compounds is essentially negligible. Table VII lists the spin- spin coupling constants as a function of mole fraction in CC14 for Pb(CI-IZCH3)4 which has an internal chemical shift of approxi- mately zero. Table VI. Proton Chemical Shifts in Ethyl Compounds Internal Chemical Internal Chemical Shifts Shifts (cps.) (cps.) At Infinite Dilution Compound Pure Compound CHC13, CCl4 C6le Zn(CHzCH3)2 -51.5 :1: 0.4 - -48.8 - Hg(CHzCH3)Z -16.2 i0.1 -18.3 '17.0 '1595 Ge(CH2CH3)4 -18.1:t 0. 2 - -18.8 - Sn(CHZCH3)4 -22. 5 :1: 0. 2 -24.0 -23.0 -22.4 Pb(CHzCH3)4 ’\‘ O - 40 - 41 42 Table VII. Spin-Spin Coupling Constants of Pb(CHZCH3)4 in CC14 —' Mole Fract1on JAB(cps.) JAX(cps.) JBX(cps.) 0.099 7.83*0.15 125.7i0.2 41.6i0.2 0.332 7.82i0.16 125.3i0.l 41.6:h0.3 1.000 7.9Zi0.22 125.3i0.2 41.0:h0.2 The value of JBX for Sn(CHzCH3)4 reported by Narasimhan and Rogers (79) is in disagreement with the value 51.4 cps. reported by Klose (90). In order to determine this coupling precisely the n.m. r. spectrum of Sn(CHzCH3)4 has been obtained at a spectrometer frequency of 100 Mc/sec. . Using an internal chemical shift of 37. 6 cps. at 100 Mc./sec. values for JAX and JBX can be obtained. These are listed in Table VIII along with the values previously reported (79). Table VIII. Spin-Spin Coupling Constants of Sn(CHZCH3)4 Compound JAB(cps.) JAX(cps.) JBX(cps.) sn“7(CHZCH3). 8. 1 73. 2 55. 7 Sn1'9(CHZCH,). 8. 1 76. 7 58. 4 Sn“7(CHzCI-l3)4a 8. 2 68. 1 30. 8 sn'"9(c112c113).f‘1 8. 2 71. 2 32. 2 3”Reference 79. 19. O . e V 18. 0 g B 17.0 e a A. .6 . hp") 16. 0- . G 1 15. 0 380493211 24- 0W 23. 0.. O V” )4 m C if ‘r "6/ V 22. O» 21. summon). -5 20.0 (0130) 19. 18. 17. 52. 51. 50. -8 . (0930) 49. 48. 0- Zamora). 47. A i I I l ' 1 1 0.0 0.10.2 0.30.4 0.5 0.6 0.7 0.8 0.91.0 Mole Fraction of Solute Figure 3. Internal Chemical Shifts Versus Mole Fraction of Solute for Ethyl Compounds. A-CC B-CHC C'GBH” 44 The n.m. r. spectra which are reproduced on the following pages were obtained using a Varian high- resolution spectrometer operating at 15. 085 Mc./sec. The spectra for each compound will be presented in pairs. For the increasing sweep direction the peaks will be point- ing up and for the decreasing sweep direction the peaks will be point— ing down. 2 Substituted Acetic Acids The chemical shifts, spin- spin coupling constants, and peak assignments for the series of substituted acetic acids are listed in Table IX. The carbon-13 n.m. r. spectra are shown in Figures 4 through 10 . T rifluoroac etic Acid The carbon-13 n. m. r. spectra of trifluoroacetic acid are shown in Figure 4. The most intense peak to the low magnetic field is due to the carbonyl carbon. The quartet at higher magnetic field is due to the trifluorom ethyl carbon. Difluorochloroac etic Acid The carbon-13 n.m. r. spectra of difluorochloroacetic acid are shown in Figure 5. The spectra consist of four peaks. The carbonyl peak occurs at lowest magnetic field. The triplet at higher magnetic field is due to the difluorochloromethyl carbon. Fluorodichloroac etic Acid The carbon-13 n.m. r. spectra of fluorodichloroacetic acid are shown in Figure 6. The spectra contain three peaks. The most in- tense peak to the low magnetic field is due to the carbonyl carbon. The doublet at higher field is due to the fluorodichloromethyl carbon. 45 meoo .Eom ofloomouosfiffi mo museum mfiuconnmo 503.32 HA: .w “:5th .mmo com I A O \ om / ommouoop I smashes“ ENS 3.1.3 5.7.0 ammo Enho meoo 46 .Eom ofloomoHofiaoosoajfic mo @30on Mascoflhmo .oom\dz HA: .m osapmflh I. QmNoHUo e EGQU 3.00 38.80 Amvsommo Amzoemo meoo 3:0...er :zofio A O )/ ommonocfi mm .mflU OON 47 .Eom ofioomouozowpouodd mo mspoomm mfiuconsmo .mmoz H .ma T 03 QMNUHUMU 3.00 5.0.3 3.3er .mmu CON K ) Eromo :Eomo smashes“ $.00 .o oudmfim 48 Trichloroac etic Acid The carbon-13 n.m. r. spectra of trichloroacetic acid in chloro- form are shown in Figure 7. The spectra contain two peaks. The peak for the carbonyl carbon occurs at low field and that for the trichloro- methyl carbon at higher field. Dichloroacetic Acid The carbon—l3 n.m. r. spectra of dichloroacetic acid are shown in Figures 8(a) and 8(b). The most intense peak at low magnetic field is due to the carbonyl carbon. The high field doublet is due to the dichloromethyl carbon. Chloroacetic Acid The carbon-13 n.m. r. spectra of Chloroacetic acid are shown in Figure 9. The most intense peak at low magnetic field is due to the carbonyl carbon. The high field triplet is due to the chloromethyl carbon. PhenLlacetic Acid The carbon-13 n.m. r. spectra of phenylacetic acid are shown in Figure 10(a) and 10(b). The spectrum consists of three groups of peaks. The carbonyl carbon occurs at lowest magnetic field. The doublet is due to the aromatic ring carbons. The triplet occurring at higher magnetic field is due to the carbon that has the phenyl group and two protons attached to it. 49 :300 0300moso30€u mo 0:033 2:02:00 003.02 Ema .N. enamfih / ommon0op m 4 0.33:9: m meoo :00 .36 com . . meoo n60 50 COZH )———-1 200 cps. CHC12(1) CHC13(2) H increase sf 0 Figure 8(a). 15. 1 Mo/sec. carbon- 13 spectrum of dichloroacetic acid. CHc1.(1) cncnm) decrease H % o COZH Figure 8(b). 15.1 Mc/sec. carbon-13 spectrum of dichloroacetic acid. 51 .300 030000H0H£0 mo 0:00am MHIGOQSU .00m\.0§ Ema .0 Headwfih meoo .mmo CON 0 Fl / osmouump 3 50:0 56.48 Exomo 23030 A a a Gem 3:060 for 30.30 Amzoemo 23030 0 E A ommouocfl $.00 .300 03000H>G0£m mo §h00omm mfiuconnso .00m\.02 Ema .Amvod ousmfih 52 o 1 033.35 E 0&0 cow N 5 no 33030 5.30 :38 :Eomo Emu :50 53 .200 030001393 mo 85:00am MHIGOQHMG .00m\.0§ H .ma mo :50 meoo .mmo OON Emu o .A mm ommohomb 5:25 3:20 5.30 £58 21:0 nHomo .32 $st Table IX. Carbon- 13 Spectra of Some Substituted Acetic Acids 54 Compound Group Peak (cps.) (ppm.) JC13_F (cps.) CF3COZH COZH 1 511 33. 94:0. 7 CF3(1) 2 818 cr3(2) 3 1106 82.8409 287:1:2 cr,(3) 4 1393 CF2(4) 5 1680 crzcmozn cozH 1 489 32. 240. 5 CFZC1(1) 2 906 297:1:4 crzcuz) 3 1202 79. 640. 4 crzcm) 4 1499 CFClzCOZH cozn 1 461 30. 540. 6 CFC12(1) 2 1179 98. 240. 9 30344 CFC12(Z) 3 1482 CC13COZH COZH 1 447 29. 63:0. 3 CC)3 2 1603 106. 240. 8 1 JC 3-H CleCOzH COZH 1 368 24. 440. 3 CC12H(1) 2 1853 128.8:I:0.4 181:1:3 CC12H(Z) 3 2034 CClHZCOZH c0211 1 338 23. 340. 7 CC1H2(1) 2 2190 CC1H2(2) 3 2345 155. 4:1:1. 0 155:1:4 OCH-12(3) 4 2500 (bonzcozn cpZH 1 236 14. 340. 3 C ,CH(1) 2 912 60640.5 CH(2) 3 1066 65. 6:1:0. 5 154i3 CHZ(1) 4 2177 CHZ(Z) 5 2313 153.440. 5 13744 0142(3) 6 2451 assig given Figu) and s fiuor shoe? Begir Thef Ineth Acen Figu: The) the c expec pounc Centr tons) quart :1: (D >4: Q) M / ShOWn 55 Some Fluorinated Aliphatic Compounds The chemical shifts, spin- spin coupling constants, and peak assignments for some fluorinated aliphatic compounds studied are given in Table X. The carbon-l3 n.m. r. spectra are shown in Figures 11 through 17. Table XI lists the carbon-13 chemical shifts and spin-spin coupling constants of some aliphatic compounds and fluorinated aliphatic c ompounds . T rifluoroac etic Anhydride The carbon-13 n.m. r. spectra of trifluoroacetic anhydride are shown in Figures 11(a) and 11(b). The spectra consist of five peaks. Beginning at low field, the first peak is assigned to the carbonyl carbon. The four remaining peaks comprise the quartet due to the trifluoro- methyl carbon. Ac etylac etone The carbon-13 n.m. r. spectra of acetylacetone are shown in Figures 12(a) and 12(b). The spectra consist of three groups of peaks. The two peaks occurring at the lower magnetic field are assigned to the carbonyl carbons. Two types of carbonyl carbon atoms are to be expected on the basis of the keto-enol equilibria existing in this com- pound. The two lines appearing at intermediate field are due to the central carbon. The doublet is to be expected since one of the two pro- tons on the central carbon takes part in the keto-enol equilibria. The quartet occurring at high field is due to the methyl carbons. Hexafluoroac etylac etone The carbon-13 n.m. r. spectra of hexafluoroacetylacetone are shown in Figures 13(a) and 13(b). The spectra consist of six peaks. 56 CF3(2) CO CF3I3) CF3(1) CF3I4) increase H > o 200 cps. Figure 11(a). 15. 1 Mc/sec. carbon-13 spectrum of trifluoroacetic anhydride. CFsil) CF3I4) CO CF30) H decrease; CF3(2) o Figure 11(b). 15.1 Mc./sec. carbon-13 spectrum of trifluoroacetic anhydride. .0Gou000H>u000 mo 93.30090 mfiuconnmd 003.02 Ema ASNH 0ndmfm 57 \ o i 0000HUGM m .mmo com 650 O0 :50 3.50 3.50 5 50 EOU 58 0:800030000 mo 593025 Mauconumo .00m\02 1m: .AQVNH 0stwm 00 $00 O K A 0mm0n00p m .mmo com 550 Samoa 550 350 3.50. :50 59 .0cou003300mouosamx0a mo Eds000mm m: #5930 .00m\ .02 H .2 40:; 0H5wfirm E50 :50 5.50 5.00 0&0 CON . O \ I 0mm0hocfi 300 E 60 .0Cou000300omou03G0x0: mo 83.3025 mausonumu .00m\.02 A .mH .3va 0Hdmfih ZOO Amvmmne AmvaHV v ai A::5o .eee oom Amvmnu «)1- .1/8 508 101 por 61 Beginning at the low field part of the spectrum, the first peak is that of the carbonyl carbon atom. The next three peaks are part of the quartet from the trifluoromethyl carbons. The fourth member of this quartet cannot be seen due to overlap with the doublet from the central carbon. The two peaks occurring on the high field portion of the spectrum are due to the central carbons. 1, 2 Dibromo- 1, 1, 2, 2, - tetrafluoroethane The carbon-13 n.m. r. spectra of 1, 2-dibromo—l, l, 2, 2, -tetra— fluoroethane is shown in Figure 14. The spectrum consists of three peaks that comprise the triplet to be expected due to the spin- spin inter- action between carbon- 13 and the bonded fluorine nuclei. 2 , 2 , 2 - T rifluoroethanol The carbon-13 n.m. r. spectra of 2, 2, 2-trifluoroethanol are shown in Figures 15(a) and 15(b). The spectra consist of two groups of peaks. The quartet occurring on the low field portion of the spectrum is due to the trifluoromethyl carbon. The triplet occurring on the high field portion of the spectrum is due to the alcoholic carbon. 1, 1, l-Trifluoro-Z- propanol The carbon-13 n.m. r. spectra of 1, 1, l-trifluoro-Z-propanol are shown in Figures 16(a) and 16 (b). The spectra consist of two quartets and a doublet. The quartet occurring at the low field portion of the spectrum is assigned to the trifluoromethyl group. The doublet is due to the alcoholic carbon. The second quartet occurring at the high field portion of the spectrum is due to the methyl carbon. 62 CF 1 CF2BI’(2) 2Br( ) 1 CF r‘ ) CFZBr( CFzBr(3) CF2B1‘(2) increase decrease ‘r H > o O 4____J 200 cps. Figure 14. 15.1 Mc./sec. carbon-13 spectra of 1, 2-dibromo-1,1, 2, 2— tetrafluoroethane. 63 .Hocmaamouosfiwuu- .N .N .N mo €350on 2-4909th .ovm\.02 1mg .3va vndmwh [.mmo CON A omMopufi sEONmo n Emommu 25098 3 mo 5P3 Esme O I 64 .Hocmsuvouosflcuum .N .N mo 8.950on Mauconnumu .oom\.o§ imd .3va 0&9me game 820 ammo EMS I, .mmo com o Acmmvhuov E Emofo EmONmu 330:5 65 .HOCMQOHQuNuouOSGCHJ J J Ho 9330QO mauconymo .uvm\.02 H .m~ 43A: 059me 5:15 @130 amen Asmomo ammo ammo Aavnmo :mmo .Cmomu :1 o Tll. A ommmnoov om .Hocmmoumumuouogflhud J J we 5.930QO mauaonnmo .uom\.02 12 4.3.: vudwwh Ede 3an Asmomu ammo Emmo O O a \ E / UvaHuGfi m ammo Asmomo Fig the the 66 1,1, 1 -Trif1uoro- Z-methyl- Z— prOpanol The carbon-13 n.m. r. spectra of 1, l, l-trifluoro-Z-methyl-Z- propanol are shown in Figures 17(a) and 17(b). The spectra consist of nine peaks. The quartet occurring at the low field portion of the spectrum is due to the trifluoromethyl carbon. The single peak at the intermediate field portion is due to the alcoholic carbon. The quartet occurring at the high field portion of the spectrum is due to the methyl carbons. a -Substituted Toluenes The chemical shifts, spin-spin coupling constants, and peak assignments for a series of a- substituted toluenes are listed in Table XII. The carbon-l3 n.m. r. spectra are shown in Figures 18 through 25. Toluene The carbon-l3 n.m. r. spectra of toluene are shown in Figures 18(a) and 18(b). The spectrum consists of two groups of peaks. The lower field peaks are those of the aromatic carbons. The quartet at high field is due to the aliphatic carbon. Benz ylbromide The carbon-13 n.m. r. spectra of benzylbromide are shown in Figures 19(a) and 19(b). The doublet at low magnetic field is due to the aromatic carbons. The triplet occurring at higher field is due to the aliphatic carbon. Benzylchloride The carbon-13 n.m. r. spectra of benzylchloride are shown in Figures 20(a) and 20(b). The aromatic carbon resonances are in the low / ‘ Suva: nu P-Nson mo 85.30696 mH uCOnHHmo .omm\ .02 H .mH .HmvHN ouderm EmmNmo :Emfio Hmvmmmmo o 4 ommoHoCH HH .mnHo OON Emu Emu .GmummonEHCHNCon Ho 5530QO mH Iconumo .oom\ .02 H .mH .SIN 634E 81 :50 Emu ) .mmu CON 0 \ 4 ommouoop TH N Emm monEwNCU 82 .ocmfiuoEHCHCoHHQHC mo 8.9.3.0on mHuconHmo .oom\.oH>H H .mH .Hmvmm madmfm Nmu \ I ommoHoCH O I Emu .646 com 1||+ SHED Mme 83 .oCCHHoEH>CoHQHC mo gnuoomm mHucooflHwo .oom\1.oH>H H .mH .HQVNN oudeh N.mo .23 com 1., o / ommohoop H.H Emu Emu U 6 £0 84 .oCH:m>oH>Nsvn Ho 5530on mHusonHmo .oom\.oH>H H.mH .7“.va oudmfim Euro :35 £38 K ll vmmoHoGH 0 mm .wmo com .llln. Emu Emu mo 85 .opHCm>oH>NCon mo 85.30on mHuconHwo .oom\.uH>H H .mH Ham... 634E. Emu Emu O ommmnuop H.H Efo 3.248 Efo 86 Benzylamine The carbon-13 n. m. r. spectra of benzylamine are shown in Figures 24(a) and 24(b). The peaks for the aromatic carbons occur in the lower field portion of the spectrum. The peak occurring at lowest field is that of the substituted carbon. The other two more intense peaks are due to the five remaining ring carbons. The triplet occurr- ing in the higher field portion of spectrum is due to the aliphatic carbon. Benzylalcohol The carbon-13 n.m. r. spectra of benzylalcohol are shown. in Figure 25. The aromatic portion of the spectrum, occurring at low field consists of three peaks. The peak at lowest field position is due to the substituted carbon. The other two more intense peaks are due to the remaining five ring carbons. The triplet occurring at high field is due to the carbon atom of the CHZOH group. Diphenyl Compounds The peak positions, observed and calculated peak heights, and peak assignments for a series of diphenyl compounds are listed in Table XIII. The chemical shifts for the various ring carbons are listed in Table XIV. The 15. 1 Mc. /sec. carbon-13 spectra for these com- pounds are shown in Figures 26 through 29. Diphenylether The carbon-13 n.m. r. spectra of diphenylether are shown in Figure 26. The spectra consist of six distinguishable peaks. The assign- ments given are based on the spectra of phenol which were analyzed in detail by Lauterbur (8) with the aid of methyl substitutions. 87 CH(l) CH(2) H increase > o C8 CHZNHZ(2) CH2 (1) HZNHZ(3) {-————4 200 cps. Figure 24(a). 15. 1 Mc./sec. carbon-13 spectrum of benzylamine. 88 CH,NH,(1) ‘CH,NH,(3) CH,NH,(2) H decreasekr o CH(l) 1—1 200 cps. CH(2) Figure 24(b). 15. 1 Mc./sec. carbon-13 spectrum of benzylamine. 89 .HoaHoonH>Nc®nH mo whuoomm mHucoonHmu .uom\.oH>H H .mH .mN oudmfim Em Emu U k 0 A vmwmhuvv Emofo :EONmo E K i omonoGH €30.35 0 TH .mmo com «I'll Emu m :Eo 90 Table XII. 15.1 Mc./sec. Carbon-13 Spectra of Some a-Substituted Toluenes V”c 6 c JCU-H Compound Peak Group (cps. ) (ppm. ) (cps .) Toluene 1 CS 706 60.140. 3 2 CH(l) 958 68.840.4 16141 3 CH(2) 1118 4 CH,(1) 2474 5 CH,(Z) 2596 176.140. 5 12241 6 CH,(3) 2718 7 CH,(4) 2839 Benzylbromide 1 CS 890 59. 040. 5 Z CH(I) 943 67. 940. 3 16241 3 CH(2) 1105 4 CH,(1) 2302 5 CI-lz(2) 2456 162. 840. 5 15443 6 CHZ(3) 2610 Benzylchloride 1 CS 894 59. 343 2 C140) 938 67. 643 16242 3 CH(2) 1100 4 CH,(1) 2125 5 CH,(2) 2275 150. 640. 5 15145 6 CHZ(3) 2426 Benzylmercaptan 1 CS 835 55.440. 2 2 CH(l) 958 68. 740. 2 15943 3 CH(2) 1116 4 CH2“) 2383 5 CH2(2) 2523 167. 240. 3 14042 6 CHZ(3) 2663 Diphenylmethane 1 CS 796 52. 842 2 CH(l) 907 65. 543 16843 3 CH(2) 1069 4 CH,(1) 2204 5 CHZ(2) 2332 154. 640. 0 13042 6 CH2(3) 2454 continued 91 Table XII - Continued Vc 6c JCl3-H Compound Peak Group (cps. ) (ppm. ) ( cps.) Benzylcyanide 1 CS 947 62. 843 2 CH(1) 947 67.940.11 15742 3 CH(2) 1104 4 CH,(1) 2488 5 CH,(Z) 2620 170.740.6 13243 6 CH,(3) 2753 Benzylamine 1 Cs 802 53. 240. 5 2 CH0) 967 69.540.5 16145 3 CH(2) 1129 4 CH,(1) 2137 5 CH,(z) 2273 150.740.8 13644 6 CH2(3) 2408 Benzylalcohol 1 Cs 841 55.740.4 2 CH(l) 963 69.1-40.5 15741 3 CH(2) 1121 4 CH,(1) 1868 5 CH2(2) 2010 133.241.0 14142 6 CH3(3) 2151 92 .HoaHuoHlaaofiHmHHu «0 0300mm mHunonHmu .0om\.0H2 H .mH 6N musmHh $ Iommouoov nu 3m .m 336 36 .N 0 HH H: 56:“. _ 56 .N H34 3.: .m :3.~ \ I an .m .mmo com o unmouoaH mm 93 Diphenyl sulfide The carbon-13 n.m. r. spectra of diphenylsulfide are shown in Figure 27. The spectra consist of two peaks indicating that the five ring carbons are equivalent. The substituted carbon overlaps the lower peak of the doublet from the ring carbons. Diphenylselenide The carbon-13 n.m. r. spectra of diphenylselenide are shown in Figure 28. The spectrum consist of four partially resolved peaks. In the assignment of the peaks advantage is taken of the fact that resonance frequencies of meta carbon atoms have a range of 2.4 ppm. (14). The 15. 1 Mc./sec. carbon-13 chemical shifts of meta carbons occur near 65 ppm. on the carbon disulfide scale. Therefore the lines giving a chemical shift near 65 ppm. relative to carbon disulfide will be assigned to the meta carbon. Diphenylmercury The carbon-13 n.m. r. spectra of diphenylmercury are shown in Figure 29. The substituted carbon resonance occurs at a lower magnetic field than the other ring carbons. T riphenyl Compounds The peak positions, observed and calculated peak heights, and peak assignments for some triphenyl compounds studied in this research are listed in Table XV. The chemical shifts for the various ring carbons are listed in Table XVI. The 15.1 Mc./sec. carbon-13 spectra for these compounds are shown in Figures 30 through 35. 94 .mvaHSmH>C0nHQHv «0 8.3.0090 2:654:00 .00m\ .02 H .mH .3. 6.33.4 Emu SEC 0 A I 0mmvnoop mmo 3:40 Emu \ 11 0mm?“ 00H 0 HIH 95 .0H0H00H00H>COHHQHH0 Ho 0300mm mHuconHHmo .0om\.0H2 12 .3 654E 4m0 0 0mmou0op v 8 .AmmmouosH 0 mm .68 cow. .11.. Nmo Em.m :34 .4 area .m Em; 5:. .m 36:“. “6.0 .H 3:50.58 100:me Ho 08.00am mHnsonHmo .00m\ .02 H .mH .om whamHh 96 £0 0 1030.000 H.H 56.2.34 Em.m SKYN 2:. 3:. 34 3.3:“. :35 \ 1’ 0m00H0CH .mm0 com .11 97 Table X111. 15. 1 Mc./sec. Carbon-13 Spectra of Some Diphenyl Compounds Height Compound Peak (cps .) Obsd. Calcd. Assignments Diphenylether 1 555 1. 70 2.00 C—1 2 894 2.15 2.00 CI-I-3, 5(1) 3 1008 0.90 1.00 . CPI-4, (1) 4 1054 4.20 4.00 CH-3,5(2);CH-2,6(l) 5 1159 1.05 1.00 CH-4(2) 6 1222 2.00 2.00 CI-l-Z, 6(2) Diphenylsulfide 1 888 6.61 7.00 C—1;CH-2,3,4,5,6(1) 2 1053 5.39 5.00 CH-2,3,4,5,6(2) Diphenylselenide 1 857 4.16 4.00 C-l;CI-l-2, 6(1) 2 905 2.71 3.00 CH-4(l);CH-3,5(1) 3 1004 1.19 2.00 CH-Z, 6(2) 4 1067 3.93 3.00 CI-I-4(2);CH-3, 5(2) Diphenylmercury 1 340 1. 73 2. 00 C-1 2 779 2. 28 2.00 CH-Z, 6(1) 3 922 4.29 5.00 CH-4(1);CH-2, 6(2); CPI-3, 5(1) 4 1072 3.70 3.00 CH-4(2);CI~I-3, 5(2) 98 Table XIV. Chemical Shifts (ppm.) and Spin-Spin Coupling Constants (cps.) For Some Diphenyl Compounds Compound Cs Ortho Meta Para Cn-H Diphenylether 36. 75.6 64.7 71. 160 Diphenylsulfide 58. 64. 2 64. 2 64. 165 Diphenylselenide 56. 61. 7 65. 3 65. 157 Diphenylmercury 22. 56. 4 66.1 66. 148 Diphenylmethane 52. 65. 5 65. 5 65. 162 99 T riphenylamine The carbon-13 n.m. r. spectra of triphenylamine are shown in Figure 30. The spectra consist of five peaks. The first peak has been assigned to the substituted carbon. The second peak is the first of the doublet due to the meta carbon. The third peak includes the first members of the doublets for the ortho and para carbons. The fourth peak is the second member of the meta carbon doublet. The fifth peak is the second member of the doublet for the ortho and para carbons. T riphenylphosphine The carbon-13 n.m. r. spectra of triphenylphosphine are shown in Figure 31. The spectra consist of four peaks. The first peak is due to the first members of the doublets from the ortho and para carbons. The second peak is due to the first member of the doublet from the meta carbons. The third peak is due to the second members of the doublets from the ortho and para carbons. The fourth peak is the second member of the doublet for the meta carbons. Triphenylarsine The carbon—13 n.m. r. spectra of triphenylarsine are shown in Figure 32. The spectra consist of four peaks with assignments as follows: The first peak is due to the first members of the doublets from the ortho and para carbons. The second peak is due to the first member of the meta carbon doublet. The third peak is due to the second members of the doublets from the ortho and para carbons. The fourth peak is due to the second member of the doublet from the meta carbons. 6558130433» mo 0:00am mHuconumo .00m\.0H>H H .mH .om 4.;th 100 .36 SN _ . 5:36 N . £38 . :FEEJ . . Ema . m To . HNMV‘LO 1A m 0m00h000 m 0 81 J 0mmohucH TH 101 .msfinamogma>G0£QHHuwo mnuoomm mHuconHmo .00m\.0v& H.mH .3 654E £0 18 I ommuu0op 0 mm .9? COM .1. RI I 33.... 36.234 Ems 36.82:. ammonocH O mm 102 .0sHmHmH>00Hmwuumo mhpoomm mH1Co3Hmo .00m\.0<4.H.mH .Nm 634; k / 0m800000 0 4H .mmu oom \ 1A Ems Staged Ema :ILSEJ 0m00H0CH 0 mm .4 .m .N .H 103 Triphenylstibine The carbon-13 n.m. r. spectra of triphenylstibine are shown in Figure 33. The spectra consist of five peaks with assignments as follows: The first peak is due to the first member of the doublet from the ortho carbons. The second peak is due to the first member of the doublet from the para carbon. The third peak is due to the first member of the doublet from the meta carbons. The fourth peak is due to the second member of the doublets from the ortho and para carbons. The fifth peak is due to the second member of the doublet from the meta carbons. T riphenylbi smuthine The carbon-13 n.m. r. spectra of triphenylbismuthine are shown in Figure 34. The spectra consist of five peaks with assignments as follows: The first peak is due to the first member of the doublet of the ortho carbons. The second peak is due to the first member of the doublet from the para carbon. The third peak is due to the first member of the doublet from the meta carbons and the second member of the doublet from the ortho carbons. The fourth peak is due to the second member of the doublet from the para carbon. The fifth peak is due to the second member of the doublet from the meta carbons. T riph enylm ethane The carbon-l3 n.m. r. spectra of triphenylmethane, aromatic portion only, are shown in Figure 35. The spectrum consists of three peaks with assignments as follows: The peak occurring at lowest magnetic field is due to the substituted carbon. The other two more intense peaks are due to the remaining five carbon atoms of the ring. The aliphatic C13 doublet was at higher field and is not shown in the figures. 104 .ocHnHumH>c0~HmHnu mo muuoomm mHLuofiumu .0om\.02 H .mH .mm 0.33% £0 \ A ommouoov O m 36 con 541m .m T1 Etfiid .4 :35 .m _ 2:. .N 36:“. .H o 1. A. coach 2: H.H 105 055953313043?» mo ouuoomm mHuconumo .000\ .02 H .mH .3 35E Nm0 K I4 ammouuov 4H .36 com Tlllli H84 Ema 34236:“. 341m :SJ N ‘ ommokunH . o-CNMV‘ID N.610 o mm .vqmfiofiacmxmwuu mo muuommm mdunonumo .umm\.o§ H .2 .mm 9:.th 106 .mmo CON 0 N / ommouoov E Nmo Emu Emu Emu o N lmmdmu a”: E Emu 107 Table XV. 15.1 Mc./sec. Carbon—13 Spectra of Some Triphenyl Compounds V: Height Compound Peak (cps. Obsd. Calcd. Assignments Triphenylamine 1 701 3. 30 3.00 C-1 2 904 3.66 3.00 CH-3, 5(1) 3 984 4.10 4.50 CH—4(1);CH-2,6(l) 4 1062 2.41 3.00 CH-3, 5(2) 5 1147 4.53 4.50 CH-4(2);CH-2,6(2) Triphenylphosphine 1 842 5.00 4. 50 CH-4(1);CH-2, 6(1) 2 907 3.85 3.00 CH-3, 5(1) 3 994 1.65 4.50 CH-4(2);CH-2,6(2) 4 1074 4.40 3.00 CH-3, 5(2) Triphenylarsine 1 824 4.45 4.50 CH-4(1);CH-2, 6(1) 2 904 4. 35 3.00 CH-3, 5(1) 3 990 1.89 4.50 (SH-4(2), CH-Z, 6(2) 4 1071 4. 30 3.00 CH-3, 5(2) Triphenylstibine 1 793 3.43 3.00 CH—Z, 6(1) 2 826 1.78 1.50 CH-4(1) 3 907 2.89 3.00 CPI-3, 5(1) 4 956 3.03 4.50 CH-Z, 6(2);CH-4(2) 5 1065 3.89 3.00 CH-3, 5(2) Triphenylbismuthine 1 768 2. 76 3.00 CH-Z, 6(1) 2 882 3.06 1.50 CH-4(l) 3 931 3.72 6.00 CH-Z, 6(2); CPI-3, 5(1) 4 1040 2. 39 1.50 CPL-4(2) 5 1074 3.06 3.00 CH-3, 5(2) Triphenylm ethane 1 754 C -l 2 906 CH-2,3,4,5,6(1) 3 1068 CH-2,3,4,5,6(2) 4 2024 CH(l) 5 2169 CH(2) 108 Table XVI. Chemical Shifts (ppm.) and Spin-Spin Coupling Constants (cps.) of Some Triphenyl Compounds Compounds CS Ortho Meta Para JCU-H Triphenylamine 46.5 70.7 65.2 70.6 150 Triphenylphosphine ---- 60.9 65. 7 60.9 156 Triphenylarsine ---- 60.1 65. 5 60.1 166 Triphenylstibine ---- 58. 0 65.4 59. 0 150 Triphenylbismuthine —--- 56. 3 66. 4 63.7 155 Triphenylmethane 50.0 65.4 65.4 65.4 162 109 Alcohols The chemical shifts for some aliphatic alcohols are listed in Table XVII. It has been empirically found by Malinowski (93) that carbon—13 proton spin- spin coupling constants, for compounds of the type CHXYZ, can be separated into three components according to the equation qu-H = 7x133?L ‘72 (56) In equation (56) J is the carbon-13 proton spin- spin coupling 13 constants; compofent:I 7X, 3Y' 32 are contributions which are associated with substituents X, Y, and Z, respectively. The additivity relation expressed in equation (56) has been derived using valence bond theory by Juan and Gutowsky (94). The valence bond approach used with this model gives a linear relation between the 5 character of the carbon hybrid orbital involved in the C-H bond (aHz ) and the observed CU-H coupling constant. = 500 a J (57) C13'H HZ Where J is the observed Cl3—H coupling constant and 0H2 is the c‘3-H fractional 5 character of the C-H bond. The experimental values for the spin- spin coupling constants J for the aliphatic carbon in the a-substituted toluenes are listed C‘3-H in Table XVIII. The calculated spin-spin coupling constant and the fractional s-bond character based on the assumption that the contact term is dominant are also listed in Table XVIII. 110 Table XVII. Chemical Shifts for Some Aliphatic Alcohols); C-O C2 C3 Alcohol (ppm.) (ppm.) (ppm.) Methyl 145 --- --- Ethyl 137 176 --- (F Iso-propyl 131 169 --—- n-Propyl 129 167 183 n-Butyl 132 158 --- :- iso-Butyl 124 162 174 sec-Butyl 125 167 175 tert-Butyl 125 167 175 :1: These data were supplied by J. Stothers (80). 111 Table XVIII. Calculated and Observed Spin-Spin Coupling Constants in a-Substituted Toluenes JCl3-H(Cps') JC13-H(Cps°) Fractional Compound Calculated Observed s-Character ¢CH3 126 122 0. 244 (bongo 127 130 o. 260 cbcnzcozn 131 137 0.274 CbCHZNI-Iz 134 136 o. 272 (bCHZCN 137 132 o. 264 cbcnzsn 139 140 o. 280 (bCHZOH 144 141 o. 282 6CHZC1 153 151 o. 302 ¢)CHZBr 153 154 o. 308 112 An attempt has been made to correlate the carbon-13 chemical shifts with the molar susceptibility of the molecules,__.YSLithin a series the change in molar susceptibility should reflect the change in bond magnetic anisotropy. The results are shown in Figures 36 through 40. Figure 36 shows the correlation of the carbon-13 chemical shifts with molar susceptibilities (81) for some aliphatic alcohols- Figures 37 and 38 show the correlation of the carbon-l3 chemical shifts with molar susceptibilities for the substituted carbon and the carbonyl carbon atoms in some substituted acetic acids, respectively. Molar susceptibility data are not available for all of the compounds studied in this research but have been estimated from Pascal constants in some cases. Figure 39 shows a linear relationship between carbon-13 chemi- cal shifts and molar susceptibilities for some phenylmethanes. Chemical shift data for tetraphenylmethane are not available due to its low solu- bility. Figure 40 shows the correlation of carbon-13 chemical shifts with molar susceptibilities for some diphenyl and triphenyl compounds. It has been shown that carbon-13 chemical shifts depend in part, on the electronegativities of the group substituted on the carbon (13, 14). The carbon-13 chemical shifts for the substituent carbons and for the carbonyl carbons in substituted acetic acids have been plotted versus the group electronegativities (82) in Figures 41 and 42, respectively. Figure 43 shows the correlation between carbon-13 chemical shifts of the aliphatic carbon atom and the group electronegativities (83) of the CHZX group in the OCHZX series. Figure 44 shows the correlation between carbon-13 chemical shifts of the ortho carbon atoms and Pauling's electronegativities (84) for the central atom of a series of diphenyl and triphenyl compounds. 113 .303002 03.3334 080m HON mowfidnwumvumdw HMHOE msmum> muflfim HMdeumSU .om 0.8:th EH N ax... 00 cm ON 00 cm ow om ON 0..” o . . . . . . 1 . . . . 2: Endgamé436624683629A J 4 :3 3585-9339694 .2 .H 8: . Hofifimouofimueé .N .N E aofimoumé £5620 E .o: Hondmonm gnawed/TN .3 3535A .3 . 3533-: 3.. 39865-». a; #1 4 u a q a .02 aondmoumua A3 . 353 3566...... E 5 .o 35:32 A: .62 n .3; 1 9: 114 200 r 190 ~ 180 ~ 170 '. 160 ' .150 — 140 - 130 _ cmrco,11 (ppm.) 120- . ‘7 n 110 " 100* 90 ~ cr,oo,n so. 1 70' 60- \ so; 40. 3O1 1 1 1 1 I I .- 0 .C W11- . 0 10 20 ‘30 40 50 60 7O 80 90 100 .110 120 ~x. 2 10° Figure 3?. OX. Chemical Shifts Versus Molarh'Susceptibilitie's in Some Substituted Acetic Acids. 115 80 - . 70- 6o. 50' 0 (ppm.) 40— 01?,ch 6 0301.00.11 ¢ 30» 01',0100,H ’5 00 o . a 1.03.8 . 01m,oo,n zo~ 0150100,}! 9 03300.11 10» - 1 l 1 a I o 20 4o - 6o 80 . 100 120 ix. x 10' Figure 38. Chemical Shifts of Carbonyl Carbon Atoms Versus Molar Susceptibilities for Some Substituted Acetic Acids. 116 200 . U 190 150 ' 140 .130- 120 ' 1 . . . 60 80 100 ,120 '140 160 180 -11. 2’10“ ~ I Figure 39. The Aliphatic Carbon Atom. Chemical Shifts Versus Molar Susceptibilities in Some Phenylmethanes. 110 117 100 1 90' 801' 70* 5o O 60* 50. 40) 30 - 100 003,89 0 . ' 1 120 140 160 , 180 ax. : 1C, '0 «p.31 ‘ . 200 220 Figure 40. Chemical Shifts of Ortho Carbons Atoms Versus Molar Susceptibilities for'Some Diphenyl and Triphenyl Compounds. . . .5: 636... 2.5664 wouaflmadm mo .mownom d HON hugfldmoaonuovflm 95.20 mdmue> ”GEM 328030 «X0 .3. onfiwmh rgwudwoaohugm 95.3” cm .m co .m cm .N 00 .N o cm 2. 2: SH Aév .0. 0mm 1.18 mpg com 1 mNN 1 omN . - th 119 . .5: 636... 6:34 363036 mo 328 a. new hugflmwoaouuoeflm macaw mdmuo> m53< £03.30 7:39.30 mo “HEW #038030 .Nv mafimfm bgflmwoconuooam 990.20 cm .m as .m om .N - co .N o 4 . 2 ado-mu _ o eased". a... 3.3 120 4me monogoh wouafiumnam . U 5 figfldwmnouuomfim macho m5mh0> 53w 33.9250 X n30 .mw onamwh , . humzudwmconuuofim @5050 cm .v mN .v co xv mu .m cm .m mm .m co .m ms .N om .N‘ mm .N 1 + . 1 A q . OCH .7 4 1 o2 . 23. 3:5 9: cm“ .. .OON. 121 m.m .mvndomEoU 1:333." .H. van 15:35 080m new 83< Hahn—~30 mo >3>3dmoaouuoo~m n.mafidmm mamno> m83< nonumo 9*qu mo mufinm 3.35050 4% uudmwm figfldwoao 503m m .mfigdnm o.m m.N o.N m4 1 3» co DISC USSION OF RESU LT S Solvent Studies of Ethyl Compounds The data in Table VI and Figure 3 show that the effects of the solvents chloroform, carbon tetrachloride, and cyclohexane on the proton n.m. r. spectra of the ethyl compounds of the metals are small. The absence of large changes in the internal chemical shifts with dilution indicates that there is essentially no intermolecular association between the solute molecules. Furthermore, there does not appear to be any appreciable intermolecular interactions between the solvent and solute molecules. The proton magnetic resonance spectra of the ethyl compounds have been analyzed and published previously (79, 85, 86). In the analysis of the proton spectra for these compounds the assump- tions are made that for a given compound the ethyl groups are equivalent and that all the methyl and methylene hydrogens are equivalent, respectively. The small differences between the internal chemicals shifts in the pure solute and at infinite dilution indicates that these assumptions are valid. Corrections of the internal chemical shifts for differences in the bulk diamagnetic susceptibility were not made due to the lack of susceptibility data for the ethyl compounds. In addition, the chemical shifts of the methyl and methylene protons were measured in the same environment with respect to solvent. The JAB’ JAX' and JBX compounds do not show any significant changes with concentration. spin-spin coupling constants for the ethyl Table VII lists the changes in these spin- spin coupling constants as a function of mole fraction of Pb(CHzCH3)4. Similar results are found for 122 123 the other ethyl compounds. Table VIII lists the values of the three coupling constants for Sn(CZH5)4 determined in this work and also the values determined previously by Narasimhan and Rogers (79). A significant disagreement occurs in the values of the JB coupling X constants. It now appears that the previously reported value for JBX is wrong and the correct value is approximately 55 cps. Correlation of Chemical Shifts with Molar Susceptibilities I Figures 36 through 40 show that the carbon-l3 chemical shifts correlate quite well with molar susceptibilities in many cases. In all i the figures with the exception of Figure 38 chemical shifts increases as the molar susceptibilities increase. The correlation of carbon-13 chemical shifts with diamagnetic susceptibility data has been predicted theoretically by Pople (17) for atoms with 2p electrons. As predicted theoretically it is observed that the magnetic moment per molecule increase as the shielding increases. Thus, Figures 36, 37, 39, and 40 all have positive slopes. However, in Figure 38 it can be seen that as the magnetic moment per mole increases the shielding decreases. This reversal of the slope for carbonyl carbons may be due to the extra paramagnetic contributions to the atomic susceptibility due to the C = 0 double bond. Whenever the applied field is oriented perpendicular to a double bond the theory predicts larger X P . terms on the atoms participating in the double bond. Correlation of Chemical Shifts with Electronegativity The correlations of carbon-13 chemical shifts with electronega- tivity for the compounds studied in this research are shown in 124 Figures 41 through 44. In general it can be seen that the carbon-l3 chemical shift of a carbon atom in a group decreases with increasing group electronegativity. Thus the shielding of a given atom decreases as the group electronegativity increases as would be expected if inductive effects were dominant. The case of the carbonyl carbons is different. InFigure 42, the carbon-13 chemical shifts of the carbonyl carbon are seen to in- crease with increasing group electronegativity of the substituted methyl group in a series of acids. It is probable that inductive effects'are not as important as magnetic anisotr0py effects here. Spiesecke and Schneider (13), in a study of the carbon-l3 resonances in compounds of the CH3CH2X type, concluded that the anisotropy contribution causes a high-field shift of the a-carbon resonance and a low-field shift of the [3- carbon resonance a similar effect may operate here. In Figure 43 the correlation of carbon-l3 chemical shift with group electronegativity is shown for some substituted toluenes, OCHZX. It seems reasonable to expect that the correlation. with electronegativity would be similar to that observed in CH3X compounds. It should be noted that ¢CHzBr, CbCHzCl, and (LCHZCN deviate from the simple electronegativity correlation line. This may be due to an appreciable neighbor anisotropy effect in these cases. Similar effects for CH3C1, CH3Br, and CH3I have been noted in correlation plots of carbon- 13 chemical shifts with electronegativity of the substituent atoms (13). Figure 44 shows a plot of ortho carbon-l3 chemical shifts in diphenyl and triphenyl compounds as a function of the electronegatiyity of the substituent atoms. As might be expected the inductive effect would be small at the ortho carbons. Hence it appears that anisotropy contribution is the dominent factor. 125 The approximate agreement of the J coupling constants c‘3—H calculated using zeta values with the observed J constants given c‘3-H in Table XVII indicates that the contact term is the dominant factor determining this coupling in the a- substituted toluenes. SUMMARY The natural abundance carbon-13 magnetic resonance spectra of several aliphatic, phenyl, diphenyl, and triphenyl compounds have been obtained at 15. 085 Mc. /sec. The spectra have been analyzed and peak assignments made. The chemical shift data show a good corre- lation with group and atomic electronegativities and molar susceptibilities. The spin- spin coupling constants have been recorded in all cases. In the d-substituted toluenes it was shown that the experimentally ob- served carbon-l3 proton spin- spin coupling constants agree quite well with the spin- spin coupling constants calculated using zeta values. Solvent studies of some metal-organic ethyl compounds were per- formed by observing the proton magnetic resonance spectra at 60. 000 Mc./sec. The solvent studies results indicated that in chloroform, carbon tetrachloride, and cyclohexane the amount of intermolecular association is negligible. The spin— spin coupling constants were ob- served not to vary with concentration. 126 10. P. 11. P. 12. P. 13. H. 14. H. 15. R. 16. M. 17. J. 18. J. BIBLIOGRAPHY . D. Cohen, N. Sheppard, and J. J. Turner, Proc. Chem. Soc., 1958, 118. . Muller and D. E. Pritchard, J. Chem. Phys., 1, 768 (1959). V. D. Tiers, J. Chem. Phys., 22, 2263 (1961). . K. Harris, J. Mol. Spectroscopy, 12, 309 (1963). C. Lauterbur, J. Chem. Phys., 26, 217 (1957). . H. Holm, J. Chem. Phys., 2_6_, 707 (1957). C. Lauterbur, J. Am. Chem. Soc., 83,, 1838 (1961). C. Lauterbur, J. Am. Chem. Soc., 83, 1846 (1961). C. Lauterbur, Ann. N. Y. Acad. Sci., 7_0, 841 (1958). C. Lauterbur, J. Chem. Phys., _3_8_, 1406 (1963). C. Lauterbur, J. Chem. Phys., 3_8_, 1415 (1963). C. Lauterbur, J. Chem. Phys., _3_8, 1432 (1963). Spiesecke and W. G. Schneider, J. Chem. Phys., _3_5, 722 (1961). Spiesecke and W. G. Schneider, J. Chem. Phys.,_3_5, 731 (1961). A. Friedel and H. L. Retcofsky, J. Am. Chem. Soc., _§_5_, 1300 (1963). Karplus and J. A. Pople, J. Chem. Phys., 3_8, 2803 (1963). A. Pople, Disc. Faraday Society, 34, 7 (1962). A. Pople, W. G. Schneider, and H. J. Bernstein, "High Resolu— tion Nuclear Magnetic Resonance, " Chapter 5, McGraw-Hill, New York, 1959. 127 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 128 W. E. Lamb, Phys. Rev.,_§_0_. 817 (1941). N. F. Ramsey, Phys. Rev.,_7_8_, 699 (1950). A. Saika and C. P. Slichter, J. Chem. Phys., 22, 26 (1954). J. A. Pople, Proc. Roy. Soc. (London), 5232, 550 (1957). H. M. McConnell, J. Chem. Phys., 21, 226 (1957). F. London, J. Phys. Radium, 8, 397 (1937). J. A. Pople, J. Chem. Phys., .2_4, 1111 (1956). E E H. Eyring, J. Walter, and G. E. Kimball, "Quantum Chemistry, " p. 99, John Wiley and Sons, New York, 1944. A. Streitweiser, Jr. , "Molecular Orbital Theory for Organic Chemists, " p. 39. John Wiley and Sons, New York, 1961. J. A. Pople, J. Chem. Phys., 17, 53 (1962). H. F. Hameka, M01. Phys., 1, 203 (1948). H. F. Hameka, J. Chem. Phys., .33, 1996 (1961). H. F. Hameka, J. Chem. Phys., _3_,'_7__, 3008 (1962). J. A. POple, J. Chem. Phys., 31, 3009 (1962). M. Karplus, and J. A. Pople, J. Chem. Phys., _3_8_, 2803 (1963). D. F. O'Reilly, J. Chem. Phys., _36, 274 (1962). C. W. Kern and W. N. Lipscomb, J. Chem. Phys., _3_Z_, 260 (1962). C. W. Kern and W. N. Lipscomb, J. Chem. Phys., _3_1, 275 (1962). H. F. Hameka, Rev. Mod. Phys., 3_4, 87 (1962). S. K. Ghosh and S..K. Sinha, J. Chem. Phys., _3_6_, 737 (1962). I. J. Maddox and R. McWeeny, J. Chem. Phys., 3_6_, 2353 (1962). C. C. Hall and A. Hardisson, Proc. Roy.Soc. (London), 268, 328 (1962). 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 129 N. Jonathan, S. Gordon, and B. P. Dailey, J. Chem. Phys., 26. 2443 (1962). H. S. Gutowsky, D. W. McCall, and C. P. Slichter, Phys. Rev., _8__2_, 748 (1951). H. S. Gutowsky, D. W. McCall, and C. P. Slichter, Phys. Rev., §_4_, 589 (1951). H. S. Gutowsky, D. W. McCall, and C. P. Slichter, J. Chem. Phys., 21, 279 (1953). E. L. Hahn and D. E. Maxwell, Phys. Rev., .8__4,, 1246 (1951). E. L. Hahn and D. E. Maxwell, Phys. Rev., _8_8_, 1070 (1952). P. L. Corio, Chem. Rev., 62, 363 (1960). P. L. Corio, and I. Wienberg, J. Chem. Phys., _3_1, 569 (1959). N. Muller, and D. E. Pritchard, J. Chem. Phys., _3_l_, 768 (1959). Reference 18, Chap. 8. M. Karplus, J. Am. Chem. Soc., 84, 2458 (1962). P. C. Lauterbur and R. J. Kurland, J. Am. Chem. Soc., _8_4_, 3405 (1962). F. A. L. Anet, J. Am. Chem. Soc., _8_4_, 3767 (1962). J. D. Graham and M. T. Rogers, J. Am. Chem. Soc., _8_4, 2249 (1962). H. S. Gutowsky and C. Juan, J. Chem. Phys., _3_7, 120 (1962). K. L. Williamson and W. S. Johnson, J. Am. Chem. Soc., _8_3_, 4623 (1961). F. A. L. Anet, J. Am. Chem. Soc., _8_‘_1_, 747 (1962). M. Barfield and D. M. Grant, J. Am. Chem. Soc., 83, 4726 (1961). 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. H. J. W. 130 J. Bernstein and N. Sheppard, J. Chem. Phys., _3_7_, 3012 (1962). . Meiboom, Ann. Rev. Phys. Chem., E, 335 (1963). M. Jackman, "Applications of Nuclear Magnetic Resonance Spectroscopy in Organic Chemistry, " Chapter 2, Pergamon Press, New York, 1959. . D. Roberts, "Nuclear Magnetic Resonance, " p. 12, McGraw-Hill, 1959. Bloch, Phys. Rev., :72, 460 (1946). . Bloembergen, E. M. Purcell, and R. V. Pound, Phys. Rev., _7_3_, 679 (1948). J. Karabatsos,J. Am. Chem. Soc., fl, 1230 (1961). . Muller, and D. E. Pritchard, J. Chem. Phys., _3_1_, 1471 (1959). K. Harris, J. M01. Spectroscopy, _12, 309 (1963). Muller, and D. T. Carr, J. Phys. Chem., 91, 112 (1963). M. McConnell, and D. D. Thompson, J. Chem. Phys., _2_6_, 958 (1957). M. McConnell, and D. D. Thompson, J. Chem. Phys., _3_1. 85 (1959). . C. Lauterbur, J. Am. Chem. Soc., _83, 1838 (1961). Bloch, Phys. Rev., 102, 104 (1956). T. Arnold, Phys. Rev., 102, 136 (1956). Freeman and D. H. Whiffen, Proc. Phys. Soc. (London), 22, 794 (1962). M. Anderson and J. D. Baldeschwieler, J. Chem. Phys., _3_Z 39 (1962). A. Anderson and R. Freeman, J. Chem. Phys., _3_1, 85 (1962). 77. 78. .79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 131 R. Freeman and W. A. Anderson, J. Chem. Phys., _3_7_, 2053 (1962). P. C. Lauterbur, in "Determination of Organic Structures by Physical Methods, " F. C. Nachod and W. D. Phillips, Ed. , p. 485, Academic Press, New York, N. Y., 1962. P. T. Narasimhan and M. T. Rogers, J. Chem. Phys., _3_4, 1049 (1961). J. B. Stothers, Private Communication. G. W. Smith, General Motors Research Laboratories, Report. GMR-317, December 20, 1960. J. Hinze, M. A. Whitehead, and H. H. Jaffe, J. Am. Chem. Soc., £5, 148 (1963). J. K. Wilmshurst, J. Chem. Phys., _2_Z, 1129 (1957). L. Pauling, "The Nature of the Chemical Bond, 3d Ed. , p. 93, Cornell University Press, Ithaca, New York, 1960. P. T. Narasimhan and M. T. Rogers, J. Am. Chem. Soc.,_§_2_, 34 (1960). P. T. Narasimhan and M. T. Rogers, J. Am. Chem. Soc., 8_2_, 5983 (1960). . B. Savitsky and K. Namikawa, J. Phys. Chem., fl, 2430 (1963). . B. Savitsky, J. Phys. Chem., 61, 2723 (1963). . B. Savitsky and K. Namikawa, J. Phys. Chem., _6_7_, 2754 (1963). G G G G. Klose, Ann. Physik, _8_, 220 (1961). G. V. D. Tiers, J. Phys. Soc. Japan, _1_5_, 354 (1960). N . Sheppard and J. J. Turner, Proc. Roy. Soc. (London), A252, 506 (1959). E. R. Malenowski, J. Am. Chem. Soc., 83, 4479 (1961). C. Juan and H. S. Gutowsky, J. Chem. Phys., 31, 2198 (1962). mmmTPY UBP‘M‘N