STUDIES OF NUCLEAR MAGNETIC RELAXATION AND MOLECULAR MOTIONS OF SOME SMALL WLECULES IN DENSE PHASES By Shaw-Guang Huang A DISSERTATION Submitted to Michigan State University in partial fulfillment Of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1977 Pulsed ‘0 per t0 the ed an: '35 de descr; PhYsi. Shift the 1 ABSTRACT STUDIES OF NUCLEAR MAGNETIC RELAXATION AND MOLECULAR MOTIONS OF SOME SMALL ‘DLECULES IN DENSE PHASES BY Shaw- Guang Huang A general program has been written to perform a large variety of pulsed NMR studies automatically. Several interfaces were constructed to permit varying the pulsewidths, to supply a constant pulsed current to the field-gradient coil and to alternate the radiofrequency pulses. 14 d 19 Complete N an F relaxation studies of CFSCN were accomplish- ed and the notional anisotropy of the molecules in the liquid state was determined. The J-diffusion model was found to be appropriate to describe the reorientational motion of the CFSCN molecules. SeVeral physical constants of the molecules were determined, e.g., the chemical 13 shift anisotropy and the spin-rotation constants for both the C and 19' the P nucleus. 13C relaxation studies have been carried out for CDBr3, CFBr3 and Cflznrz and the scalar coupling constants between 13C and 798r determined from the dominant scalar relaxation rates in these highly brominated compounds. In CHZBr2 the coupling ConStant d(C793r) was fOund to be 58.7 Hz, a reasonable value compared with the reported value 1 to be I discuss lined : test 01 cussed. Au CF f‘r‘ “by n 4 4 degree liquid perform dipolar lotion 1 relax”: “but fl NMR lin nuclei. IOIatin diffusi j“? t: orders lattice Shaw~Guang Huang value for J(C35Cl). But,in both CDBr and CFBr3,J(C7gBr) was found 3 to be about twice as big as expected and this discrepancy has.been discussed. The coupling constant J(F798r) in CFBr3 was also deter- mined and found to be 30.5 Hz from the 19F relaxation studies. The test of the extended diffusion model in this molecule has been dis- cussed. An attempt at resolving the anisotropic motions of planar CFZCCIZ molecules in the liquid state has been made with a certain degree of success.It was found that the reorientational motion in the liquid and in the solid states was via different mechanisms. The separation of various 19 F relaxation mechanisms in CF31 was performed but difficulties were found in extracting the intramolecular dipolar relaxation rates and therefore analysis of the reorientational motion was fruitless. However, it was found that the spin-rotational relaxation mechanism was via the spinning motion of the molecules . about the C3 axis. A number of plastic crystals were studied by analysis of the 19? NMR lineshapes and by measurement of the relaxation rates of the 19F nuclei. Measurement of the 19F spin—lattice relaxatiOn rate in the rotating frame was found to be very informative about the self- diffusive motion in the plastic phase of these crystals. The mean jump time between adjacent diffusion steps was determined to several orders of magnitude. In all the plastic crystals studied the Torrey lattice-diffusion model was found to be appropriate to describe the self-diffusion of the molecules in the plastic solid phase. The reorientational motion in these plastic crystals was found to be as fast as in the liquid phases. From the 19F spin-spin relaxation rates, the seal: derivati‘ deteni; e in terms Shaw-Guang Huang the scalar-coupled relaxation rates in someof'these fluorochlorocarbon derivatives were extracted and the coupling constants J(F35Cl) determined. Finally, a test of the theory of fusion was discussed in terms of the activation parameters derived from the NMR studies. To my wife and my parents ii ”'1 In encoura He also some of Depart: author. h'ZRYTIe B: ACKNOWLEDGEMENT The author wisheS‘tothank Professor Max T. Rogers for his encouragement and freedom tOlPursue his own interesting projects. He also wishes to thank Dr. David A. Wright for initial guidance in some of the instrumental work. Appreciation also goes to the Department of Chemistry for offering teaching assistantships to the author. Kindly help from Mr. Martin Rabb, Mr. Frank Bennis and Mr. Wayne Burkhardt is also highly appreciated. iii TABLE OF CONTENTS LIST OF TABLES . . . . . . LIST OF FIGURES. INTRODUCTION . . . . . . THEORIES OF MOLECULAR MOTIONS IN DENSE PHASES. I. II. III. IV. Introduction . Correlation Functions and Transport Coefficients . A. Velocity Auto-Correlation Function. 8. Angular Velocity Correlation Function . C. Angular Momentum Correlation Function . MOdels of Molecules in Dense Phases. A. Hard Sphere MOdels. . 8. Potential Functions and the Self-Diffusion Coefficients. . . . . . . . . . . . Models of MOlecular Motions in Dense Phases. A. Debye's Model of Brownian Motion. 8. Free Rotors . . . . . . . . . . ; . . . . C. Extended Diffusion Models . Theories of Fusion . . . . A. Melting Transition from the Instability of the Liquid and Solid Phases . . . . B. Coexistence Theory. c. Lattice Models. . . . . . . . . . . . D. Extended Lennard-Jones-Devonshire Models. iv . xiii . 10 . 13 . 13 . 16 . 17 . 18 . 22 . 23 l. Pople and Karasz's Theory of Fusion. 2. Amzel and Becka's Modification . THEORIES OF NUCLEAR MAGNETIC RELAXATION . I. Historical Review. . A. The Bloch Equation and the Origin of Magnetic Relaxation. . B. Density Matrix Formalism for Magnetic Relaxation. II. Mechanisms of Nuclear Magnetic Relaxation. A. Dipolar Relaxation. B. Quadrupolar Relaxation. C. Scalar Relaxation . D. Chemical Shielding Anisotropy Relaxation. E. Spin-Rotation Relaxation. . . . . . . . . . F. NMR Relaxation due to Anisotropic Reorientation . III. Relaxation and Fourier Transform NMR Spectroscopy. A. Theories in Fourier Transform NMR Spectroscopy. B. Fourier Transform Nuclear Magnetic Relaxation Studies . . . . . . C. Effects of Cross-Relaxation and Cross-Correlation . EXPERIMENTAL. . . . . . . . I. The NMR Spectrometer . . . . . . . . . . A. General Description of the Pulse System . B. Multinuclear Modification . C. The Probes. . . . D. Computer Interfaces . E. External Lock System. . . . . . . . . . . . . F. Temperature Control and Measurement . . . V 23 26 28 28 28 30 32 33 36 37 38 38 41' 43 43 45 47 51 $1 51 54 55 59 62 62 G. Heteronuclear Decoupling and the Variable Field Lock System . . . . . . . . . . . . . . . . . . . . 63 II. Measurement of Relaxation Times. .. . . . . . . . . . . 67 A. NMRLX-- A General Program for Computer Controlled Pulsed NMR Relaxation Studies . . . . . . . . . . . 67 B. Sample Preparation. . . . . . . . . . . . . . . . . 68 RESULTS AND DISCUSSION. . . . . . . . . . . . . . . . . . . . . . 70 I. Molecular Motion and Relaxation Studies of Some Small Molecules in the Liquid Phase. . . . . . . . . . . . . . 70 A. Anisotropic Reorientation and 13C, 14N, 19F Relaxation Studies in CF3CN . . . . . . . . . . . . 70 1. Introduction . . . . . . . . . . . . . . . . . 70 2. 14N Relaxation and the Tumbling Motions. . . . 71 3. 19F Relaxation and Anisotropic Reorientation.. 74 4. The Extended Diffusion Model . . . . . . . . . 90 S. 13C Relaxation and the Spin-Rotation Constant. 92 6. 13c, 19F Chemical Shifts and JCF . . . . . . . 97 B. 13C and 2D Relaxation Studies and the Motional Anisotropy in Liquid CDBr3. . . . . . . . . . . . . 100 1., Introduction . . . . . . . . . . . . . . . . . 100 2. 2D Relaxation and the Tumbling Motion. . . . . 103 3. 13C Relaxation and Motional Anisotropy . . . . 106 c. 19F and 13c Relaxation Studies in Liquid CFBr3. . . 114 1. Introduction . . . . . . . . . . . . . . . . . 114 2. 19F Spin-Lattice Relaxation and IsotrOpic ReOrientation. . . . . . . . . . . . . . . . . 114 3. 19F Spin-Spin Relaxation and J(F7gBr). . . . . 121 vi II. F. 4. 5. 13C Nuclear Magnetic Relaxation and J(C79Br) in 13 C Spin-Lattice Relaxation and J(C798r) Extended Diffusion Models. Liquid CH23I2 . 1. 2. 3. Introduction . Mechanisms 0 Di f 13C Relaxation and J(C798r). scussion . Anisotropic Reorientation and 19F Relaxation in CFZCClz. l. 2. 3. Introduction . Mo Ro lecular Motions in the Solid Phase . tational Motion in the Liquid Phase. 19F Relaxation Mechanisms for CF31. NMR Studies of Molecular MOtions in Plastic Crystalline Fluorochlorocarbons. A. B. C. Introduction. NMR Theories Relating to Molecular Motions in Plastic Crystals . l. 2. NMR Studies of Some Fluorochloroethanes CanCl6_n.. 1. Li newidths . . . . . . Second Moments of NMR Absorption Lines . Spin-Lattice Relaxation Times. Spin-Lattice Relaxation Times in the Rotating Frame . . . . . . . MO a. lecular MOtions in CFZClCC13 . Measurement of the 19F NMR Linewidths and Second MOments . vii . 124 . 125 . 131 . 131 . 133 . 136 . 138 . 138 . 139 . 145 . 147 . 152 . 152 . 154 . 154 . 156 . 158 . 160 162 . 163 . 163 b. Rotational Motion in the Plastic Phase . c. Self-Diffusion in the Plastic Phase, d. 19F Relaxation in the Liquid Phase and J(F35c1) 2. Molecular Motions in CFCIZCFC12 . a. Measurement of the 19F NMR Linewidths and Second Moments . b. Rotational Motion in the Plastic Phase . c. Self-Diffusion in the Plastic Phase. d. 19F Relaxation in the Liquid Phase and JCF35C1) 3. Self-Diffusion Studies in CF3CC13 . 4. Spin-Lattice Relaxation Studies in CF2C1CF2C1.. D. Molecular Motions in Some Other Plastic Crystals . l. l,l,l-Trifluoropentachloropropane (TFPCP) 2. 2,2,4,4-Tetrachlorohexafluorobutane (TCHFB) 3. l,1,2,2-Tetrachloro-3,3,4,4-tetraf1uoro- cyclobutane (TCTFCB). E. Discussion . . . . . . . . . . . . . . 1. Rotational Motion in Plastic Crystals . 2. Self-Diffusion in Plastic Crystals. 3. Test of the Theory of Fusion. . . . . . SUMMARY . . . . . 1 REFERENCES. APPENDIX. . A. Instructions for NMRLX - A General Program for Computer Controlled Pulsed NMR Studies. viii . 167 . 173 . 178 . 184 . 184 . 187 . 190 . 195 . 195 202 . 206 . 206 . 211 . 217 . 226 . 226 . 228 . 229 . 231 . 233 . 248 . 248 II. III. IV. VI. VII. VIII. IX. X. XI. XII. INTRODUCTION. CONSTRUCTION OF THE PROGRAM . LEGAL COMMANDS. TIMING. DISPLAY . WEIGHTED LEAST SQUARES CURVE FITTING. SETUPS FOR EXPERIMENTS. (i) Pulse Adjustment. (ii) T1 Measurement Using the Spin-Inversion- Recovery Technique. (iii) T1 Measurement Using the Homospoil Technique . (iv) Tlp Measurement . (v) T2 Measurement. (vi) Th Measurement. (vii) Dsd Measurement . USEFUL SUBROUTINES. ERROR MESSAGES. AUXILIARY PROGRAMS. USEFUL PATCHES. INTERFACING CIRCUITS. . Instructions for Other Programs Used . I. II. III. FTNMRD. FTNMRB. SECMFT. ix 248 248 249 252 254 255 257 257 259 260 261 261 262 265 266 267 268 269 271 273 273 273 274 TABLE 10 11 12 13 14 15 16 17 18 LIST OF TABLES Some experimental measurements and the related time correlation functions . Pulse methods for relaxation studies . 14N spin-lattice relaxation rates in CF3CN . 19F spin-lattice relaxation rates in CF3CN . Measurements of the density of CF3CN . Self-diffusion coefficients of CFSCN . Anisotropic motional parameters for CFSCN. Comparison of some physical constants and motional parameters of acetonitrile and the trihaloacetonitriles. Reduced reorientational and angular momentum correlation times for CFSCN. Spin-rotation constants of 13C in CF3CN. 13C and 19 Comparison of F chemical shifts and JCF values in some fluorocarbon derivatives. Summary of the physical constants and the motional parameters of CF3CN. 2D spin-lattice relaxation rates in CDBr3. 13C spin-lattice relaxation rates in CDBr3 . The separation of the different contributions to the total relaxation rates of 13C in CHBr3 . Physical properties of CDBr3 and CDC13 . l9F spin-lattice relaxation rates in CFBr3 . Comparison of the theoretical and experimental values Page 48 72 76 80 81 83 84 89 96 99 . 102 . 104 . 107 . 112 . 113 . 115 of the intermolecular dipolar relaxation rates for CFBr3.118 TABLE Page 19 19 F spin-spin relaxation rates in CFBrS . . . . . . . . . 122 20 Summary of the 13 C spin-lattice relaxation rates in CFBr3 . . . . . . . . . . . . . . . . . . . . . . . . . . 126 21 Summary of the information used in the calculation of T1(7gBr) and J(c793r) in CFBr3. . . . . . . . . . . . . . 128 22 Physical properties of CFBr3 used or Obtained in this study. . . . . . . . . . . . . . . . . . . . . . . . 130- 23 Summary of the quantities employed in the calculation of the 13C dipolar and scalar-coupled relaxation rates in CHZBr2 . . . . . . . . . . . . . . . . . . . . . 134 24 Summary of quantities used in the calculation of T1(F79Br) and J(c793r) for CHZBrz . . . . . . . . . . . . 137 25 19F spin-lattice relaxation rates in CF2=CC12 . . . . . . 140 26 19F spin-spin relaxation rates in CF2=CC12. . . . . . . . 141 27 19F spin-lattice relaxation rates in CF31 . . . . . . . . 148 28 Self-diffusion coefficients of CF31 . . . . . . . . . . . 149 29 Separation of the contributions to the 19 F spin-lattice relaxation rates in CF31. . . . . . . . . . . . . . . . . 150 30 Physical constants of some fluorochlorocarbons. . . . . . 164 31 19F linewidths and second moments in CFZCICC13. . . . . . 16S 32 19F spin-lattice relaxation rates in CFZCICC13. . . . . . 168 33 19F spin-lattice relaxation rates in the rotating frame in CFZClCCl3. . . . . . . . . . . . . . . . . . . . 174 34 19F spin-spin relaxation rates in CF2C1CC13 . . . . . . . 175 35 Summary of the calculation of the angular momentum and the reorientational correlation times in CFZCICClS. . . . 181 xi G.- (I! 46 47 48 49 A1 TABLE Page 36 19F linewidths and second moments in CFClZCFC12. . . . . 18S 37 19F spin-lattice relaxation rates in CFCIZCFCIZ. . . . . 188 38 19F spin-lattice relaxation rates in the rotating frame in CFCIZCFCIZ. . . . . . . . . . . . . . . . . g . 191 39 19F-spin-spin relaxation rates in CFCIZCFCIZ . . . . . . 192 40 19F linewidths and second moments in CFSCC13 . . . . . . 197 41 19F spin-lattice relaxation rates in the rotating frame in CF CCl 3 3 . . . . . 198 1 42 9F spin-lattice relaxation rates in CF2C1CF2C1. . . . . 203 43 19F linewidths and second moments in CFSCIZCCCIS . . . . 207 44 19F spin-lattice relaxation rates in CF3C12CCC13 . . . . 208 45 19F spin-lattice relaxation rates, linewidths and second moments in CF3C12CCC12CF3 . . . . . . . . . . . . 213 46 19F spin-lattice relaxation rates, linewidths and second moments in 1,1,2,2-tetrachloro-3,3,4,4- tetrafluorocyclobutane . . . . . . . . . . . . . . . . . 218 47 19F spin-lattice relaxation rates in the rotating frame in l,1,2,2-tetrachloro-3,3,4,4-tetraf1uorocyclobutane... 219 . 48 19F spin-spin relaxation rates in l,1,2,2-tetrachloro- 3,3,4,4-tetrafluorocyclobutane . . . . . . . . . . . . . 220 49 Motional parameters of some plastic crystals.. . . . . . 227 50 Comparison of the parameters F and v in some plastic crystals . . . . . . . . . . . . . . . . . . . . 230 Al Discrete times and possible errors in the NMRLX timing routine . . . . . . . . . . . . . . . . . . . . . 253 xii FIGURE LIST OF FIGURES Page Plot of reduced reorientation correlation time versus reduced angular momentum correlation time for various diffusion models; solid lines A and C are predicted by the M-diffusion model for 2:1 and 2, respectively; solid lines B and D are predicted by the J-diffusion model for i=1 and 2, respectively; dashed lines are predicted by the Debye classical diffusion model (TGTJ=I/£(£+1)kT); dotted lines are predicted by the perturbed free-rotor model (Te=TJ/2£+1) and broken lines are the free-rotor limit of the extended diffusion model (T6=TJ/22). . . . . . . . . . . . . . . . . 19 Block diagram of the pulse system for 19F observation ...... 52 Block diagram of the multinuclear pulse system. . . . . . . S3 Field-gradient coils in the probe: (a) anti- Helmholtz configuration; (b) quadrupole configuration ..... 56 The inside cross-section of the wideband probe: a, sample tube; b, spinner; c, BNC connector for the capacitor pad; d, H1 tuning capacitor; e, heater sensor connector; f, N2 gas inlet; g, H1 rf cable; h, heater sensor; i, copper-constantan thermocouple; j, decoupling coil; k, sample coil; 1, external lock probe; m, decoupler tuning capacitor; n, LC circuit of the probe ..... 58 Schematic diagram of the pulsed field-gradient current supply (resistors are given in ohms, capacitors are given in uF).. . . . . . . . . . . . . . . . . . . . . . . 60 xiii HOPE 10 ll FIGURE Page 7 Schematic diagram of the rf phase reversing circuit (all resistors are in ohms, all capacitors are in uF, unless denoted otherwise). . . . . . . . . . . . . . . 61 8 (a) Circuit diagram of the decoupler gating box (all resistors are in ohms, all capacitors are in uF). (b) Timing diagram of the outputs . . . . . . . . . . . . 64 9 Block diagram of the heteronuclear decoupling setup. . . . 65 10 Block diagram of the variable field lock system. . . . . . 66 11 Design of the NMR sample tube . . . . . . . . . . . . . . 69 12 Plot of the 14N spin-lattice relaxation rate in CF3CN as a function of reciprocal temperature. . . . . . . . . . 73 13 Reorientation correlation times and diffusion coeffi- cients of CFSCN as functions of reciprocal temperature.... 75 14 Plot of the 19 F spin-lattice relaxation rate in CF3CN as a function of reciprocal temperature. . . . . . . . . . 77 15 Plots of X'I’ xi for liquid CFSCN. For comparison, the dashed lines are the corresponding X plots for CH3CN and dotted lines are those for CC13CN. . . . . . . . . . . 91 16 Plot of reduced reorientational correlation time versus reduced angular momentum correlation time for CF3CN; solid and Open circles are results with and without correction for Ceffe respectively. . . . . . . . . . . . . 93 17 Plot of JCF versus 5F for some fluorocarbon derivatives. . . . . . . . . . . . . . . . . . . . . . . . 98 xiv FIGURE 18 19 20 21 22 23 24 25 . Page Plot of AOF versus dF for CF3X, X= CC13, F, Cl, Br, I and CN; solid circles are in the C-F bond coordinate system and Open circles are in the molecular-axis coordinate system . . . . . . . . . . . . 101 Plot of the 2D spin-lattice relaxation rate in CDBr3 as a function of reciprocal temperature. . . . . . . . . 105 13 Plot of the C spin-lattice relaxation rate as a function of reciprocal temperature. The open circles are the Ric values calculated from the 13C data of 03 Farrar gt_al,2 ; the solid circles are the RE values for 13C in CDBr3. . . . . . . . . . . . . . . . . 108 Plot of the 19F spin-lattice relaxation rate as a function of reciprocal temperature; 0: Rt 1,F‘ . d . . 8r . __. R1,F, ---. RI’F. . . . . . . . . . . . . . . . . . . 116 Plot of the reorientational correlation times as functions of reciprocal temperature for CFBrS; ———-: Te; ---: TE . . . . . . . . . . . . . . . . . . . 120 Plot of the 19F spin-spin relaxation rate in CFBr3 . . t as a function of reelprocal temperature; 0: R2 F ; 3 o: R56 . . . . . . . . . . . . . . . . . . . . . . . . 123 2,F° Plot of the 13C spin-lattice relaxation rate in CFBrS as a function of reciprocal temperature; SC 1,C........ Plot of reduced reorientational correlation time 0: Rt -———- Rd 1,C; . 1,C; ---: R . 127 versus reduced angular momentum correlation time forCFBr3;O:15;0:TE.................129 XV FIGURE 26 27 28 29 30 31 32 33 Page Plot of the 13C spin-lattice relaxation rate in CHZBrZ as a function of reciprocal temperature . . . . . 135 Plot of the 19F relaxation rate for CF2=CC12 as a function of reciprocal temperature . . . . . . . . . . 142 Plot of the 19F spin-lattice relaxation rate in CF31 as a function of reciprocal temperature . . . . . . 151 19F NMR linewidths (scale at left) and second moments (scale at right) of CFZClCCIS. . . . . . . . . . 166 Plot of the 19F relaxation rate in CF2C1CC13 as a function of reciprocal temperature; : R}; 0: R5; ---- aft; —: Rd; ---: Rsr; R3“. . 17o Plot of the 19F spin-lattice relaxation rate in the rotating frame at three different spin-locking fields and the spin-spin relaxation rate (___J in.CF2ClCCl3.... 176 Plot of the mean jump time for self-diffusion in the plastic phase of CFzClCCIS; o: from Rlp data; 0: from the linewidth data using Equation (162); 0: from the linewidth data using Equation (163); o: from the linewidth data using Equation (164); +: from R2 data.... 177 Plot of reduced reorientational correlatiOn time versus reduced angular momentum correlation time for CFZCICCIS; solid lines A and C represent the theoretical curves predicted from the M-diffusion model for 2:1 and 2, respectively; solid lines B and 0 represent the theoretical curves predicted from the J-diffusion model for i=1 and 2, respectively; FIGURE 34 35 36 37 38 39 Page dashed lines represent the theoretical result predicted from the Debye classical diffusion model (TeTJ=I/£(L+1)kT); dotted lines represent the theoretical result predicted from the per- turbed free-rotor model (Te=TJ/2£+l) and the broken lines represent the free-rotor limit of the extended diffusion model (T9=TJ/22). . . . . . . . . . . 182 19F NMR linewidth (scale at left) and second moment (scale at right) for CFClZCFClz . . . . . . . . . 186 19F spin-lattice relaxation rate (0) and the spin-spin relaxation rate (0) in CFCIZCFClz; the stright solid line represents the intramolecular dipolar relaxation rate R?; -«-: the scalar spin-spin relaxation rate Ric. . . . . . . . . . . . . . . . . . . 189 Plot of the 19F spin-lattice relaxation rate in the rotating frame at three spin-locking fields (10.5, 5.3 and 2.0 G) and the spin-spin relaxation rate (0) in the plastic solid phase of CFCIZCFCIZ. . . . 193 The mean jump time derived from R10 (0), R2 (0) and Av (A) of 19 F nucleus in the plastic solid phase of CFCIZCFCIZ. . . . . . . . . . . . . . . . . . . 194 Plot of the 19F NMR linewidth and second moment in CFSCCl3 versus reciprocal temperature. . . . . . . . . . 199 Plots of the 19F spin-lattice relaxation rate in the rotating frame at three different spin-locking fields (17.6, 8.8 and 5.7 G), and of the spin-spin xvii FIGURE 40 41 42 43 44 45 46 47 relaxation rate (0) versus reciprocal temperature, for CF3CC13 . Mean jump time for self-diffusion in the plastic phase of CF3CC13 derived from Rlp(0), R2(+) and Av(0). Plot of the 19F spin-lattice relaxation rate in CFZClCFZCl versus reciprocal temperature; the solid straight line represents the intramolecular dipolar relaxation rate and the dotted line represents the intermolecular dipolar relaxation rate in the solid phase. Plots of the 19F linewidths and the second moments in CF3C12CCC13 versus reciprocal temperature . Plot of the 19F spin-lattice relaxation rate in CF3C12CCC13 versus reciprocal temperature; dashed line represents the intramolecular dipolar relaxation rate . Plot of the 19F NMR linewidths and second moments in CF3C12CCC12CF3 versus reciprocal temperature. Plot of the 19 F spin-lattice relaxation rate in TCHFB versus reciprocal temperature. Plot of the 19F NMR linewidths and second moments in C61 CF CF CCl versus L________J 2 2 2 19F spin-lattice relaxation rates in the rotating frame at three different spin—locking fields (12, 6.3, and 3.2 G) and the spin-spin relaxation rate (0) in CCI CF CF CCl L________J 2 2 2 2 2 . reciprocal temperature , xviii Page . 200 . 201 . 204 - . 209 . 210 . 214 . 215 . 221 . 222 FIGURE 48 49 A1 A2 Page 19F spin-lattice relaxation rate (0) and spin-spin relaxation rate (A) in TCTFCB; solid circles are RZ-Rl; dashed line represents the intramolecular dipolar relaxation rate; dotted line represents the total spin-lattice relaxation rate after subtracting the intermolecular dipolar contribution . . . . . . . . . . . . . . . . . . 223 The mean jump time for self-diffusion in the plastic phase of TCTFCB derived from R10 (0), Ri (0) and Au (A). . . . . . . . . . . . . . . . . . . . 224 Pulse sequence used in NMRLX . . . . . . . . . . . . . . 258 Decoupler gating box circuit diagram. (Capacities are in uF and resistances are in O). . . . . . . . . . . 272 xix INTRODUCTION Ever since the first nuclear magnetic resonance (NMR) experiments were performed, the technique has been a powerful tool in the study of the prOperties of molecules. The extreme sensitivity of the behavior of nuclear spins to their chemical and physical environment has made their detection not only useful for elucidating the detailed structure of the molecules but also the interactions among the molecular aggregates. Molecular motions, which have been studied over a long period of time, can range from being very simple as in dilute gases to being highly complex and intractable as in dense fluids and solids. While the other spectrosc0pic methods can also give information about molecular motions, NMR studies possess several advantages in deriving motional parameters such as the anisotrOpic reorientational diffusion constants and the angular momentum correlation times which cannot be obtained in most other spectroscopic studies. The success of NMR studies of molecular motion lies in the measurement of magnetic spin relaxation times which-give an estimate of the extent of perturbation by the neighboring spins through various kinds of mechanisms. Very often, there are more unknowns than relation— ships which can be obtained from the experiments. Therefore, in this 3 CF3Br, CFBr3, CDBrs and other pseudo-spherical molecules forming plastic work only simple molecules with higher symmetry,e.g., CF CN, CF I, CF C1 CCCIS, CF3C12CCC12CF3 2’ 3 2 and CFZCCIZCCIZCF2 were studied. The study of motions in these molecules crystals, CFSCClS, CF ClCCl 2 3, CFCIZCFCI was mainly to test the validity of various models such as the anisotropic reorientational diffusion model and the extended diffusion model. In additior sphhro. these TE heat ha amount BO fictions Theoret s aatEr . ful- ('9 : (n of III: 2 addition, some other information like the scalar-coupling constants, spin-rotation constants, chemical shift anisotrOpy and the relaxation times of quadrupolar nuclei was Obtained. The increasing application of minicomputers to control the experi- ment has made the measurements more rapid and accurate. A considerable amount of effort has been devoted to this as described in the Experi- mental section of this dissertation. Both historical and recently updated theories about molecular motions and magnetic relaxation are reviewed and-outlined in the Theoretical section. The working equations are then derived for use later in the data analyses in the Results and Discussion sections. As the plastic phase in solids has been useful as a bridge for further understanding of the melting processes, detailed description Of its behavior is attempted in the last part of this work. THEORIES 0F IDLECULAR LDTIONS IN DENSE PHASES I. Introduction Interest in molecular motions in dense phases has been increasing over the years. Numerous studies, both theoretical and experimental, have been devoted to describing the microdynamic behavior of molecules in liquids and solids. Unlike the gaseous state in which the mean free path is large compared to the size of each molecule, the intermolecular interaction in dense fluids where molecules are very close to each other, is dominated by a repulsive fbrce caused by frequent collisions. To describe the motion of molecules in such a dense phase, one has to deal with a many-body problem which can only be attacked by using statistical mechanical approaches with certain approximations. At present, the theory of molecular motion in dense phases is at a model-building level. Recently, irreversible statistical mechanics has been developed to supply two approaches to study hydrodynamic properties. The first one is the derivation of general master equations from which the states of thermodynamic equilibrium, hydrodynamic equations, transport coefficients, fluctuation and relaxation phenomena may be deduced as appropriate solutions to the master equationslez. The second approach is the derivation of time correlation function expressions for transport coefficients and the relaxation parameters. Both the master equations and time correlation fOrmulas can, in prin- ciple, provide solutions to kinetic and transport prOperties of many- body systems if some potential models, mathematical approximations and other techniques, e.g., cluster diagrams, are introduced. Since NMR and other spectroscopic experiments all study the 3 respo: theory functj and t1 aaCIOS E06161: £Unctj EXist: 4 response of a system coupled with an applied weak field, linear response theory has been found to be successful in describing the experimental results. The major conclusions of this theory can be stated as follows: Whenever two systems are weakly coupled to each other,such as radiation weakly coupled to matter, or molecular vibrations weakly coupled to molecular motion, it is only necessary to know how both systems behave in the absence of the coupling in order to describe the way in which one system responds to the other. Furthermore, the response of one system to the other is completely describable in terms of time correlation functions of dynamic properties. ~ In this chapter, we first introduce the time correlation functions and their relations to the transport coefficients which describe the macrosc0pic behavior of molecules. Then we introduce some molecular models, each with its characteristic shape, along with the potential functions, and motional models usually adopted. In the last section, the existing theories of liquid-solid transitions are described. II. Correlation Functions and Transport Coefficients Time-dependent correlation functions have been familiar for a long time in the theory of noise and stochastic processes. In recent years, they have become very useful in many areas of statistical physics and 'spectroscopy3’4. Correlation functions provide a concise method for expressing the degree to which two dynamical properties are correlated over a period of time. For two dynamical properties A and B, the correlation function CAB(t) is defined as CAB(t) = (1) 5 where the bracket indicates an equilibrium ensemble average. When A and B are different properties, C is called a cross-correlation function. AB When they are identical, CAB is called an auto-correlation function.‘ Since the response of a system to a specific weak perturbation field is directly related to the correlation function, many experiments have been devised to determine the specific correlation functions. For example, IR lineshapes can be used to determine the auto-correlation function of a unit vector along the molecular axis, neutron scattering experiments can determine the Van Hove scattering function? which is related to the number density auto-correlation function, and Raman scattering can determine the auto-correlation function of the second- order Legendre polynomial of the direction cosine of a unit vector along the molecular axis. Table 1 shows some familiar correlation functions and their relations to the experimental observables. The calculation of time correlation functions might be as difficult as the calculation of a partition function of an equilibrium system because the complex many-body problem cannot be bypassed. While the Monte Carlo technique6 has been applied to evaluate the multi- dimensional integrals in the solution of equilibrium partition functions, a similiar type of computer experiment, namely, molecular dynamics, has been constructed by Alder and Wainwright7, who have been able to calculate the number-density correlation function of hard-sphere and square-well 9 used a Lennard-Jones molecules in fluids. Later, Rahman8 and Verlet (12-6) potential function to simulate liquid argon and found it to be very successful. More work of this type,and the general molecular dynamics calculation of correlation functions, has been reviewed by Berne and Harplo. Tatl correlation functions. Table 1. Some experimental measurements and the related time Experimental observable Time correlation function Dynamic parameter Self-diffusion coefficient Rotational diffusion coefficient IR lineshape Raman lineshape NMR spin-rotation relaxation time <3(O)-3(t)> <fi(O)-fi(t)> <3(O)-3(t)> + m 0: linear velocrty of the mass center. 0: angular velocity about the molecular axis. 0: unit vector on the molecular dipolar axis. A U. 3: angular momentum about the mass center. \'\‘ rotatit tiles . VEIOCi magnit Scalar then q 7 NMR relaxation times are related to the self-diffusion coefficients, rotational diffhsion coefficients and angular momentum correlation times which, in turn, are related to proper correlation functions as discussed in the following sections. A Velocity Auto-Correlation Function If a molecule has a velocity 0(0) at time t=0 and a velocity u(t) at time t, then a suitable measure of the degree of correlation of these velocities is the scalar product u(0)~ u(t). After a sufficient time, the magnitude and direction of u(t) will no longer be related to u(0) and the scalar product will be zero. The velocity auto-correlation function is then defined by sw(t) = /. (2) For a collision-free gaseous molecule, the velocity is almost constant at all values of t, hence sw(t)=1. In a perfect Einstein solid in which a molecule vibrates about its site at constant angular frequency m5, sMt): cos wt. In a liquid, sw(t) has a shape of a damped oscillator. Each correlation function has its Fourier transform or its spectrum sMm) SW») = [L] sectl-e‘iwtdt. (3) 2n w It is easy to show that sMm) shows a peak at m=0 for ideal gases and a peak at m=wB for ideal solids. In liquids,sw(w) is spread over a wide frequency. We then define a diffusion coefficient Ds as the zero- of t? at: COIT and The . 8 zero-frequency intercept of 5Wm) or, in another words, the time average of the velocity correlation function, i.e., D = 5M0) J Sw(t)dt. (4) 5 on The two limits of D5 are 0 (for ideal solid) and co(for dilute gas). B. Angular Velocity Correlation Function If a molecule is rotating about its axis with angular velocity 9(0) at time t=0 and a velocity 0(t) at time t, then the angular velocity . . . 1‘ . correlation function rMt) and its spectrum V(w) are defined as rwtt) = / (5) and r‘Hw) = 1 fro(t)-e‘i‘”tdt. (6) r3; .. The rotational diffusion coefficient is then defined as' D = 1[31(0) =f rlp(t)dt. (7) r G r . Both w(t) and Dr have the same limits as for 5Mt) and 03° C. Angular Mbmentum Correlation Function If J(O) is the angular momentum of the molecule at time t=0 and it is changed to J(t) at time t by some forces, e.g., collisions of the a*‘ 3153 mOdel Para: that 3 ft the int. 9 neighboring molecules, then the angular momentum correlation function and its spectrum can be defined by jwctl = / (8) and j‘l’(w) = 1 f j111(t)°e-imtd1:. (9) f2? .. In dilute gases, collisions are rare and J is kept constant for all time, hence jw(t)=1, and jMm) peaks at m=0. In perfect solids, there is no rotation, hence J(t)EO and jw(t)50. There is no transport coefficient defined for this correlation function. To further calculate these time correlation functions, a specific model has to be chosen in order to describe the dynamic behavior of the parameters concerned. This will be discussed in the following sections. III. Models of Molecules in Dense Phases A. Hard Sphere Models The first difficulty of statistical mechanical calculations is the unknown potential function which is so complicated in dense phases that the exact expression is always impossible. In order to construct a feasible potential function, people start to build models to represent the shape and the physical properties of real molecules from which the interaction is determined. Very often, molecules in liquids are considered to be a number of spE conseq deDen; A e" there will e surfa: L - the,‘ I; 10 of spheres interacting with each other by frequent collisions. The consequence of a collision is the changing of the velocities which again depends on both the inner and surface properties of contacting spheres. For simplicity, the spheres are usually considered to be hard,i.e., there is no elasticity allowed. In other words, two colliding spheres will experience an infinite potential wall upon contacting. As for the surface properties, three different models have been popular in the theoretical derivations: 1. Smooth Spheres: The surface of the spheres is considered to be completely smooth, there is no friction exerted by contacting and therefore the tangential velocity of each colliding sphere is invariant. 2. Rough Spheres: The surface of the spheres is considered to be completely rough, therefore two colliding spheres will experience a reversal of their surface velocities at the point of contact. 3. Loaded Spheres: The center of mass of the sphere does not coincide with its geometric center, therefore a collision will cause a transfer of translational and rotational energies. B. Potential Functions and the Self-Diffusion Coefficients We now turn back to consider the appropriate potential function which can describe the interaction among the hard sphere molecules. The simplest one is called the impulse potential which is infinity when spheres are in contact and is zero otherwise, or V(r..) = , (10) lJ 0 r..>o 11 wheretris the hard-core diameter and rij is the intersphere distance This impulse potential has been used to calculate the self-diffusion , 11,12 coefficient and obtain 0 = 3 (it) (11) 8nozg(o) flm for equivalent spheres in a fluid. Here, n is the number density, m is the mass and g(O) is the contact value of the pair-distribution function. Another potential function, which is called the square-well pot- ential, has been considered to be more realistic and still allows one retain simplicity in calculationls. This is defined as fellows: V(rij) = -e O =fff(Oo)f(O) °P(QO,O)P(O,t/Oo)d00dfl, (24) she Eons“ time b'he‘, 15 where P(90) is the initial distribution probability. For example, if f(Q) is the unit vector u(t), which is related to 0(0) by 0(0) u(t)=cos 6(t), where 6(t) is the angle between 6(0) and fi(t), the correlation function can be written as C(t) = = )rcose-P(0)P(6,t)d0. (25) Since the initial distribution probability is unity, using Equation (23) for P(6,t), i.e., P(0,t)= Y1’0(0) e'ZDrt, Equation (25) gives <fi(0)-fi(t)> = = exp(-2Drt). (26) A similiar approach can be used to obtain = <%(3c056-1)> = expc-oort), (27) where P2(x) is the second-order Legendre polynomial of the argument x. In general, if the dynamic parameter can be written in terms of the Legendre polynomial Pn(cose), the correlation function can always be described by an exponential function or a linear combination of exponential functions with time constants Te=(l(2+1)Dr)’1. If one calculates the time correlation function of the angular momentum of the molecule, one finds that the angular momentum correlation time TJ is related to Teby I T T = O J 6kT ' (28) where I is the moment of inertia of the molecule. 16 Several assumptions made in this model, e.g., the infinite mass of the melecules, stick boundary conditions or assumption that the media are continuous, etc., have limited its applicability to certain systems. There have been many effbrts devoted to modifying or generalizing this 17 classical treatment of the molecular rotation. Brot and Darmon have proposed a model fer a rotor undergoing nearest neighbor jumps in an eight-wall model of cubic symmetry. Cukier and Lakatos-Lindenberg18 have treated this diffusion model by allowing arbitrary size for the random jumps and found the same exponential decaying shape fer the time correlation functions. They also generalized the stochastic model by considering the effect of finite duration of collisionslg. Berne20 has modified Debye's relaxation time by considering the effect of long-range interactions. An excellent review of the stochastic models of molecular reorientation has recently be given by SteeleZI. B. Free Rotors The other extreme of describing molecular reorientation is the gas-like free-rotor model. In this model, the molecules are depicted as undergoing essentially free rOtation, governed by their moments of inertia, but occassionally interrupted by a collision. The resulting time correlation functions can be calculated by Equation (24) and the relation between the reorientational correlation time and the angular momentum correlation time can be described by ‘3 t6 . m e (29) 21,30 Steele has considered this model but used a Gaussian fUnction 17 as the short-time approximation of the correlation function to alleviate the physically unrealistic long-time behavior of the exact free-rotor model applied to dense phases. Atkins31 has also considered this gas-like model using a quantum mechanical formalism and obtained a result which has Steele's result as its classical limit. C . Extended Diffusion Models Both the classical diffusion model and the free-rotor model have been found to suffer several defects. A more adequate theory which is capable of representing the smooth transition from free rotation to highly hindered rotation has been constructed first by Gordon22 and has 23-28. In this model, several been extended by several other authors assumptions are made. First, the molecules are assumed to undergo a succession of impulse collisions with infinitesimal collision time. Second, the molecules are assumed to rotate freely during the collision ,With a time much longer than the collision time. Third, during a collis ion, molecular orientations are unchanged but the direction of the angu1ar momentum vector and/or its magnitude is randomized after the con-ision. It is called J-diffusion if both the direction and the magnitude 0f the angular momentum are changed by a collision. If only the direction is randomized by the collision, it is called M-diffusion. A The original work of Gordon only dealt with linear molecules. The a13Plication to both spherical-top and syumetric-top molecules has _ been done by McClungzs. The approach from the memory function f0rmatlism has been given by Berne and Harp10 and also by Bliot and Constant26-28 29 . McClung has given an updated review with discussions of Mn . “Y experimental results . Of Sl.‘ fro: beCC 18 The most striking success of this model can be viewed from a plot * of T6, the reduced reorientational correlation time defined by 'k . . . Te= Tel/I7kT ,where I is the moment of inertia of the molecule, versus * TJ For a comparison, the plot of Tgversus T , the reduced angular momentum correlation time defined by T3= TJ/JI/kT. * J described by various models is shown in Figure l. for a spherical-top molecule V. Theories of Fusion Although the melting point has been widely used to characterize substances, the detailed dynamics of the melting process has never been well understood. This is again because of the unsolvable many—body problem encountered in describing the dynamic behavior of molecules in dense phases. Up to now, only a limited amount of work on this subject has been reported and the derived theories are still at a primitive stage. Nevertheless, we shall review some of these efforts, especially those which have been rather successfully applied to the study of plastic crystals. A. MeltingTransition from the Instability of the Liquid and Solid Phases This is a two-pronged attack on the problem of freezing. Approaching from the solid, an old theory of Lindemann32 predicts that a solid becomes unstable with respect to melting when the amplitude of vibration of the molecules is about one-tenth of the lattice spacing. From the liquid side, Alder and Wainwright7 have calculated the pair-correlation function of a hard-sphere liquid and found that a long-range oscillatory 19 Figure 1. Plot of reduced reorientational correlation time versus reduced angular momentum correlation time for various diffusion models; solid lines A and C are predicted by the M-diffusion model for 2:1 and 2, respectively; solid lines B and D are predicted by the J-diffusion model for i=1 and 2, respectively; dashed lines are predicted by the Debye classical diffusion model (TBTJ=I/£(£+1)kT); dotted lines are predicted by the perturbed free-rotor model (Te=TJ/22+l) and broken lines are the free-rotor limit of the extended diffusion model (Te=TJ/2£). 21 character was evident when the volume of the liquid molecules becomes 1.65 times of that for closest packing. This has been considered as a signal of freezing. Brot33 has also shown that including the attractive part of the intermolecular potential will bring a drastic amplification of the density fluctuation of a hard-sphere liquid. In 34, both molecular dynamic and recent studies of Lennard-Jones systems Percus-Yevick calculations show the same long-range oscillatory character of the pair-correlation function at kT/c= 0.72 and p03: 0.85 (p is the density and e and O are two parameters in the Lennard-Jones potential function, see Equation (18)). B. Coexistence Theory Jancovici35 has proposed a theory in which one can use different type of approximations for the description of liquid and solid phases and produce separate theoretical expressions for the free energy, A, then locate the equilibrium pressure of two states by a double tangent method, i.e., set _ aAsolid = _ 8Aliguid . av T av T where Tm is the temperture at equilibrium and V is the volume. This theory has been adopted by Barker and Henderson who have constructed an equation of state for the liquid and used the cell theory of solids to produce a melting line for argon with fair success36 22 C. Lattice Models There have been some simple lattice models which restrict molecules to the sites of a two or three dimensional lattice and exhibit a freezing transition when one of the two interwoven sublattices ofwhich the lattice is made becomes preferentially occupied. For example, in the Lennard- Jones and Devonshire mode137, a solid has its particles centered on the sites of a regular 1attece(o sites) which is interlaced with another set of sites(B sites). At zero temperature, the particles are all centered on 0 sites representing a perfect order and a liquid is repre- sented by a perfect disorder which has both a and 8 sites equally occupied. All the thermodynamic functions can then be written in terms of a parameter which is related to the energies of on and OB occupancies. It has been found37 that the Lennard-Jones-Devonshire model agrees amazingly well with many features of the experimental data for spherical or near-spherical molecules. Several variations of this simple site model have also been . considered. Runnels38 has found that if multiple occupancy of the sites and also simultanuous occupancy of nearest neighboring sites are dis- allowed, the model shows a second-order transition from an ordered to a disordered state. Bellemanssg’4 0 has also found that by adding another assumption, i.e., the second and the third neighboring sites cannot be simultanuously occupied, the model will show a first-order solid-liquid phase transition. The success of the Lennard-Jones-Devonshire model has been found to break down when it is applied to molecular crystals which do not follow the law of corresponding states(which is a consequence of the Lennard-Jones 6-12 potential field). Two extensions of this model, 23 which take into account the possibility of different orientations for the molecules, have been attempted and will be discussed in the following section. D. Extended Lennard-Jones-Devonshire Models The failure of the Lennard-Jones-Devonshire theory when applied to molecular crystals is because the intermolecular forces often deviate appreciably from spherical form and the possibility of hindrance to ro- tation in the solid and liquid states is not taken into account. A remarkable modification of this model has been given by Pople and Karasz41 , and later by Amzel and Becka42 , who considered order-disorder phenomena involving both orientation and positions of molecules and developed a theory which can successfully predict the thermodynamic properties of melting and of solid-solid transitions. Since their theories have been widely employed in the NMR studies of plastic crystals, an outline of the derivation of their equations is given here. 1. Pople and Karasz's Theory of Fusion Pople and Karasz41 modified the Lennard-Jones-Devonshire model by allowing molecules with two orientations on either'an a or 8 site so that four possibilities, al, a , 81 and 82 are to be considered. By defining 2 W as the energy difference between a and the neighboring 8 site, W' as the energy difference between two orientations on each site, 2 as the number of neighboring 8(or o) sites surrounding each o(or 8) site and 2' as the number of the closest equivalent sites, the partition . 4 . function 3 can be written as 24 N N Z = f O = f exp(-(N W + N W' + N W')/kT), (30) where NOB’ N and NB 8 are the number Of pairs associated with o o energies W and W', f is the partition function per molecule in a state of perfect order and is a function only of molecular volume and temperature, and the summation is over all orientations as well as all arrangements of particles on sites. Let Q be the fraction of molecules on a sites and S be the fraction of molecules in orientation 1, then N , N and N can be replaced GB alaz 8182 by Q, S, N, z and 2'. Hence O = :2: y(Q,S) exp(-{zNWQ(l-Q)+z'WNS(l-S)(l-2Q+2Q2)}/kT), (31) Q.3 where y(Q,S) is the number of ways to arrange particles for given Q and S, i.e., Y(Q.S) = ( N’ )2- (NQ)’ (NQ)!(N(1-Q))! (NQS)!(NQ(1-S))! 1- u . (N( 0)). . (32) (N(1-Q)S)!(N(1-Q)(1-S))! 43 Using Stirling's theorem , we can maximize (1/N)ln O to obtain Q = ( z" - z'"' 5(1-S))(2Q-1) (33) l-Q 2kT kT and 25 S z'W' l-S kT 1n (1-2Q+2Q2)(25-l). (34) By solving for Q and S, the order-disorder of molecules can be well defined. From Equations (33) and (34), we can see that Q and S are only dependent on two parameters, K=ZW/kT and v=z'W'/zW, as are the other thermodynamic functions which can be calculated from the partition function43. Several interesting results can be obtained after calculating the pressure, reduced temperature of melting, volume change, entropy of melting, etc: (1) When v<0.325, two separate transitions may be distinguished. One occurs when S=1/2, corresponding to orientational disorder, the other when Q=l/2, corresponding to positional disorder or melting; the S=1/2 transition occurs at a lower temperature. (2) When v30.325, only one transition, with relatively larger entropy and volume change than case (1), is predicted. (3) When v21, two transitions are observed again, but the S=l/2 transition occurs at a higher temperature than the Q=1/2 transition, does. These results successfully predict the characteristics of plastic crystals (case (1)), brittle crystals (case(2)) and liquid crystals (case (3)). The parameter v, which is the ratio of z'W' to zW, can resonably be assigned a physical meaning, i.e., the ratio of the reorientational energy barrier to the translational energy barrier of the mOlecules in a crystal lattice. 26 2. Amzel and Becka's Modification 4 . . . Amzel and Beckz 2 have conSidered an even more realistic case. 1Fhey assume the number of orientations on each site can be more than two and can be represented by another adjustable parameter D. Using the :5mme statistical approach, Equations (33) and (34) were modified to ‘loecome Q 2 w z'W' osi 1n = ( - 31(1- ——)) (212-1) (35) l-Q 2kT kT 20-2 23nd Si z'W' 031 2 ln = - ln (D-l) - (l- )(l-2Q+2Q ), (36) 1-81 kT D-l ‘where Si is the fraction of molecules with orientation i. The introduction of the parameter D leads to a few more features for this model: (1) When v<0.325, the orientational disorder (when S=1/D) occurs Ioefore the positional disorder (when Q=1/2), and with a constant v, a (larger D will shift the solid-solid transition, where the orientational (disorder occurs, to occur at a lower temperature but will not change ‘the melting point. (2) The threshold v value for the existence of a solid-solid 'transition increases when D increases, e.g., v=0.325 if D=2 and v=0.5 when D=60 . (3) When D>2, both the solid-solid and the melting transitions become first order . (4) The entropy of a solid-solid transition increases with D and v. 27 (5) The relative volume change on melting becomes smaller than that at a solid-solid transition when v>0.28 and D>2. The number of orientations D for a molecule is dependent on its symmetry. For example, D=6 for molecules with Td symmetry, D=16 for those with C3v symmetry and D=20 for those with O3}" symmetry, etc. It is then possible to predict the thermodynamic properties of a molecule with a knowledge of v and its symmetry. In NMR studies, the energy barriers for both the reorientational and the translational motions can be determined with a reasonable accuracy and therefore one should be able to test the validity of this theory. This will be discussed later in our studies of plastic crystals. THEORIES OF NUCLEAR MAGNETIC RELAXATION I. Historical Review A - The Bloch Equation and the origin of Magnetic Relaxation After the first discovery of nuclear magnetic resonance by Purcell Eiarmd his coworkers44, and independently by Bloch and his coworkers45, a huge number of studies, both theoretical and experimental, have been undertaken to develop the subject. In the earliest theoretical studies 0f NMR relaxation, time-dependent perturbation theory has often been used to derive phenomenological equations of the macroscopic spin lrlzmgnetization. Consider a system composed of N spin-8 nuclei placed in a Si1teady magnetic field which splits the energy into two states a and B. At thermal equilibrium, the spin populations in the on and 8 states (Na and NB) follow Boltzman's law, N __a_= exp(AE/k'r) = 1+A_E.+l(AE_)2+ see... (37) NB kT 2 kT ‘Vllere AB is the energy gap between the two levels and the second equality is from Taylor's expansion of exp(AE/kT) . Equation (37) implies that NO is bigger than NB at thermal equilibrium and there is a net magnetization M along the direction of the applied field. If a random time-dependent Perturbation field V(t) is applied to this system, a transition between tho energy levels occurs and therefore the populations are changed. To see the effect of V(t), time-dependent perturbation theory can be used which gives the transition probability46 PoB as 28 29 POLB = 12] exp(-AEt/M)dt , (38) n vulnere the long bar represents an emsemble average. The effect on the sslpin populations can be expressed by dNa “—dt ‘ ”sped Napae; (39) if we let Na: N+n and N3: N-n, then dn _ _ "‘00 52'- -n(PaB+PBa) + N(PBa-Pa8) - — T1 , (40) ‘efliere T1: l/(PaB+P8a) and no: N(p8o'Po8)/(POB+PBO)' Equation(40) shows the irleet populations of the spin system and therefore the net magnetization ‘\r:mries according to the first-order kinetic equation with a time constant .1P21’ which is called the spin-lattice relaxation time. Bloch47 has considered a classical mechanical description of the net 1:<)tal magnetization, which is represented by a vector M. In a steady Unzignetic field R0, along the z direction, M will precess about 2 due to a 1t<>rque produced by Ho. The motion of'M can then be described by — = (N x no). (41) E3)? replacing n with the net magnetization Mz (Mz=yhn), Equation (40) I> = Trpr(t)}. (46) Consider a spin system embedded in a bath(lattice) formed by its ssaarrounding molecules, the total Hamiltonian can be written as JC = 3132+ 3c + J-CI, (47) L Where JCz is the Zeeman term, JCL is the lattice Hamiltonian and JCI includes all of the interactions between spins and lattice. By realizing that the lattice has very many degrees of freedom and its energy levels are almost continuous so that it can be treated classically, and also that the coupling between the spin system and the lattice can be assumed to be Small so that the spin density operator o(t) can be obtained by tracing 0 (t) over the lattice degrees of freedom, one can obtain the equation of ulotion of the spin density matrix element50 32 aaa.(t) = -i[Jc,,o(t)]aa. -23 (t), (48) R 7 IO l 88! an 88 88 makiich has been used as a master equation for the derivation of NMR ‘Jreelaxation theories. Here, ROO'BB' are called the elements of the re- laxation matrix (tetradic) and are given by Roo'BB' = Jaee'e'(wee) * JOBa'B'(wo'B') " Ga'B'g: JYBYGWYG) ' 6018; JYG'YB'(wa'Y)’ (49) ‘eflmere _ . in. . JGG'BB'(w) fodr e Gaa'BB'(T) (50) Elrid Goo'881 = (“IJCI(O)|“'>° (51) ‘VViLth a specific interaction Hamiltonian JCI, the correlation function G <=Eln be calculated by choosing an appropriate model of molecular motion 6111d the relaxation matrix elements can be obtained. II. Mechanisms of Nuclear Magnetic Relaxation As we have discussed, any time-dependent fluctuating local field Elround the spin system will contribute to the relaxation of the spins. V'G3 outline here some of these which are important to nuclear magnetic 33 sp in relaxation . [\H. Dipolar Relaxation The Hamiltonian for dipolar interaction between two spins I1 and I2 can be written as51 2 = + ch " Ylen T1°U°Iza (52) inrlnere U is the dipolar coupling tensor. It is traceless and can be (clziagonalized by a coordinate transformation. The three principal 1<2<3mponents Of the diagonalized dipolar coupling tensor are U1 = (1-3cosze)/r3, (53) U2 = (sinecosee'mflr3 (54) Elrid u3 = (sinZGe-21¢)/r3, (55) ‘Vliere r, 6 and O are the three spherical polar coordinates of the inter- Illiclear vector in a space-fixed coordinate system. Since T1 is related 1t<3 the transition probability, which has a form similar to Equation (38), We rewrite Equation (38) as p (a) = Ll|2f e'iwtdt, (56) OB n2 34 where A is the time-independent part of JCd and V is the time-dependent part of JCd. If Il=Iz=k, then the spin system has three energy levels for four- 1 I: _ 1 2 (“182+“281 where the last two states are degenerate. It is easy to show51 that the eigenstates Yl=a1a2, w-1=B182’ To: spin-lattice relaxation rate R1 of this spin system is R1 = P1,O(mo) + 2P1,-1(2“o)’ (57) where “O is the energy gap between adjacent levels in unitsof M and P1,0, P1”1 can be calculated by Equation (56). The integral in Equation (56) is the Fourier transform of a correlation function of the time-dependent part of JCd, or the power spectrum. Assuming that the molecules undergo an isotropic Brownian motion with r, the internuclear distance, fixed, R can be calculated to be52 l 3 4 2 T 4T R1 =——*"( c + ° ). (38) t 10 6 1+ 2 2 1+4 2 2 ro r onc more where Tc is the correlation time of the motion. If r is also time- dependent, it should be included in the correlation function and the evaluation of R1 becomes more difficult. Bloembergen, Purcell and Pound52 have considered an isotropic diffusion model and found that R1 which deScribed the intermolecular dipolar relaxation can be expressed by 4 2 311 R1 - IO'Y :0”, (59) tran 35 where N is the number of molecules per cm3, a is the molecular radius and D is the self-diffusion coeficient. Torrey53 has used a random-walk model in which the molecules undergo a random flight of distance 2 and found that nT R1 = 11.292y4h21(1+1) —‘—1- , (60) tran 2 where n is the number of spins per unit volume and Td is the mean jump time and is related to the self-diffusion coefficient by 2 _ gr > , (61) where r is the jump distance. AbragamS4 considered also an isotropic diffusion model but obtained 4 2 R1 - m . (62) tran 503 HertzSS reconsidered Torrey's random-walk model and derived an expression NT ‘ 2 d (1+ 2a tran a3 SDTd 4 R = 3ny4M21(I+1) 1 )- ' (63) As we can see these equations all have the same dependence on a, D and N, but the numerical factors are all different. In the case of two nonequivalent spins, a factor of 2/3 should be introducedss, that is, in the extreme narrowing limit, 36 YZYZKZ his = 4—2—— S(S+1)Tc (64) r0t rIS and 2 2 Y R15 = - i 5(5+1) (RII ). (65) 3 2 1 tran YI tran When the molecular motion is not isotropic, then Tc should be 56 . . replaced by f(Q,D) , where Q=6,¢ are the Euler angles of the interacting vector with respect to the molecular axis system and D is the diffusion tensor. This will be discussed later. B. Quadrupolar Relaxation Nuclei with spin 8 usually have a quadrupole moment which can react with the local electric field gradient. The interaction Hamiltonian is JCQ = I 6(t) 1, (66) where Q is a tensor with elements proportional to the quadrupole 2 moment Q and the electric field gradients BZV/axz, aZV/ayz, a V/axay, 54 etc. Using density matrix formalism, we can obtain , in the limit of << more 1, 3 21+3 2 2 2 R = -{ -———-———9(1+n l3)(e Q/M) T . (67) 1Q 40 12(21_1) qz c where qZ is the 2 component of the electric field gradient tensor and 37 n=(q -q )/q is the asymmetry parameter of the electric field gradient. x y 2 C. Scalar Relaxation In the study of proton relaxation times in aqueous solutions of ++ +++ . - Mn and Gd ions, Bloembergen found that TzfiTl and is dependent on the observing frequency. This was recognized as due to the strong scalar coupling interaction between proton and metal ions. The Hamiltonian can be written as JC = I°A°S (68) . . . 54 land it contributes to the relaxation terms 2 T 2A 1,S ‘ R1 = -—-3 S(S+1)———2 (69) sc 1+A(l)IST1,S and 2 T A R =_s(S+1)(T 4 1,3 ) (70) 2 3 1.8 “A 2 T2 1 5° “13 1,5 where A is the scalar coupling constant in radians/sec, T1 S is the relaxation time of spin 8 and AmIS is the difference between the resonance frequencies of spin I and spin 8. Spin S may be the total spin quantum number for one or more unpaired electrons. It may also be the spin of a fast relaxing nucleus with a large coupling constant (e.g., Br nuclei in SnBr4S7) which will also contribute a considerable scalar coupling term to the relaxation rate. 38 D. Chemical Shielding Anisotropy Relaxation Chemical shielding is not necessarily isotropic but may be represented by a tensor; the three principal components 01,02 and O3 can be different and therefore give a time-dependent modulation of the Zeeman energy 3c = -1h(l-O)HOI + tho-81-I, (71) where O is the average chemical shielding and :1 is the anisotropic part of the shielding tensor. If 3' has three principal values Oi, 05 and O' the contribution to R1 will be51 3, 2T R - fie +O'2+O'2) , (72) 1csa 10 1 2 l+m2T2 o c . = . . . 5 if 0 is axially symmetric then 2 2 2 Tc R1 = Ig-wo(Ao) 2 2 (73a) csa 1+w T 0 C and ' 6T R = — w2(A0)2(8T + ——c—-). (73b) 2 9O 0 c 2 csa l+onc E. Spin-Rotation Relaxation 52 In the study of proton relaxation in H2 gas, Bloemgergen et al. found that the spin-rotation interaction makes a large contribution to 39 R1. This is given by R = 2Y2(%H'ZJ(J+1) + 3H"2 J(J+1) (74) lsr (zJ-I)(2J+3) Tc’ where H' is the field at one proton due to molecular rotation and H" is the dipolar field due to another protons. The first discovery of spin-rotation relaxation in liquids was by Gutowsky and coworkers58 who found that several Freon's had shorter relaxation times for 19F than for 1H and a different temperature dependence. They attributed this behavior to a greater spin-rotation interaction in the case of the fluorine nuclei and gave an expression for Tl,sr in a linear molecule. Johnson and Waughsg’6o also estimated that there is a significant spin-rotation relaxation contribution in liquid methane. A few years later, Hubbard61 successfully derived a profound theory of spin-rotation relaxation for a spherical-top molecule undergoing an isotropic Brownian reorientational motion and a Langevin type of motion for the angular velocity. Using standard density matrix formalism, he obtained ZIkT 2 T1 2 2 T12 R1 = ———§'((2CL+C”) -——7—§'+2(C“-QL)'_—7f77—- (75) SI 9” 1+on1 1+on12 where C , C” are components of the spin-rotation tensor, T1 is the -1 -1 -1 ‘ . 12-Tl +T2 , here T2 is the reorientational correlation time and I is the moment of inertia. In the angular momentum correlation time and T limit of onl’ m T <>2A . If H1 is a 900 pulse, then 7H1tp=T/2 where tp is the pulse width.‘Therefore the pulse width has to be short enough such that tp << 1/4A. (90) In an actual case, a series of pulses are applied to the system and the free induction decays (FID's) are coherently cumulated to increase the signal-to-noise (S/N) ratio. If the pulse interval is represented by T, then the pulse train will contain numerous sidebands of the rf carrier frequency w, and these will be l/T Hz apart. This pulse interval T determines the density of the sideband and therefore the resolution and the faithfulness of the representation of the spectrum. Since Fourier transform NMR is a computer assisted experiment, some theories of sampling72 have to be considered. First of all, in order to perform a Fourier transform, the transient FID has to be digitized as discrete data points. From information theory, a sine wave must be sampled in time at least twice during each cycle to be properly repre- sented. Therefore , the sampling rate should be at least twice the spectrum width to be observed. Second, since the sidebands of rf pulses are equally distributed on both positive and negative sides of the carrier frequency, but with opposite phases, the resulting spectrum will have possible back-folded peaks if only one phase detector is used. The development of the quadrature phase detecting technique74“80 and the use of crystal filters have been able to alleviate this problem. Third, the frequencybdependent phase error due to the time lag of detection has to be considered. Other theories of computation in Fourier 45 transform NMR have been discussed by Cooper69. B. Fourier Transform Nuclear Magnetic Relaxation Studies The most striking advantage of the Fourier transform NMR technique is the rapidity of data collection which has made possible the studies of nonequilibrium states of nuclear spin systemsgl. Measurement of relaxation times in a spin system containing more than one closely- spaced transition was very difficult before the Fourier transform NMR spectroscopy was developed. Vold §£_§1382 had first succeeded in measuring the spin-lattice relaxation time of a multiline system using a pulse Fourier transform NMR technique. It was found thereafter that many other pulse methods can also be combined with the Fourier transform technique to study nuclear spin relaxation. The initial work of Vold 35_ .51, using a nonselective 180o pulse to invert the system into a state of negative spin temperature, and then observing the recovery to equili- brium by means of Fourier transforming the FID's generated at different times, is called the inversion-recovery technique (IRFT). The transformed spectra are called partially relaxed spectra in which each transition follows a recovery equation of the form Mi = Moi(1-2exp(-t/Tli)), (91) where M1 is the amplitude of the magnetization of spin i and T1. is its 1 characteristic spin-lattice relaxation time. Markley EE_§1383 have devised a saturation recovery technique (SRFT) in which the system is brought into a state of infinite spin 46 temperature by a 90° pulse, then the off-diagonal elements of the spin density matrix are destroyed by violating the magnetic field homogeneity and the recovery of each spin is again detected by observing the FID's generated at .different times which follow the equation Mi = MOi(1-exp(-t/T1i). (92) Freeman and Hill84 have also devised a progressive saturation technique (PSFT) to observe the dynamic equilibrium reached by means of a repetitive pulse sequence. The cumulated equilibrium magnetizations have amplitudes which depend on the interval between adjacent pulses in the sequence and follow Equation (92) . Many other new techniques have been developed to utilize the great time saving advantage of Fourier transform NMR. Canet, Levy and Peat85 InOdified Vold's inversion-recovery technique to. develop a fast inversion- IVectovery Fourier transform (FIRFT) method in which the long time delay between each pulse sequence is shorter. Xaptein gtfl.86 also modified the IRFT method by changing the observing pulse to 30° in order to do a sing 1e-scan T1 measurement (SSFT) . Gupta87 developed a variable nutation method in which T1 can be obtained from the dependence of signal ampli- tude upon the nutation angle of the pulse. Other methods”.93 are all balSed on different arrangements of the pulse sequence. Levy and Peat have reviewed several different techniques for T1 measurement of 13C in organic compounds. The Fourier transform technique has also been applied to determine the spin-spin relaxation time of each individual peak of a complex . SPeCtrumgs. This method (called Fourier transform spin-echo spectroscopy) 120,121 miilizes a Carr-Purcell pulse sequence to generate a spin-echo 47 train and Fourier transforms the FID following the last echo maximum96'98. However, this method cannot be used in case there is homonuclear coupling since it tends to modulate the spin echoesgs. A number of 96-106 on the effects of this modulation of spin echoes by studies homonuclear coupling, and sometimes by the effects of chemical exchange, have been reported. An excellent review has been given by Freeman and Hill107 recently. Several authors have constructed single-scan pulse sequences to 92,93 9 measure both T1 and T at the same time . Edzes 2 has summarized 2 these methods and given a detailed error analysis. Table 2 gives a summary of the pulse sequences used in relaxation studies. C. Effects of Cross-Relaxation and Cross-Correlation The existence of two or more nonequivalent nuclear spins very often 108 first observed that in complicates the relaxation studies. Solomon hydrofluoric acid the recovery curve of the longitudinal component of the magnetization was nonexponential due to the cross-relaxation effect caused by the other nonresonant spin. Noggle109 also analyzed the cross-relaxation effect of coupled homonuclear spins in 2,3-dibromo- thiophene. In the relaxation studies of both proton and fluorine nuclei in fluoroform, Hubbard110 found that cross-correlation also caused a nonexponential decaying of the magnetization. Werbelow and Marshall111 extended Hubbard's calculation and applied it to both two-spin systems and to the more common cases which have methyl or trifluromethyl groups attached to a large molecule. Campbell and Freeman112 have reviewed the influence of cross-relaxation on NMR spin-lattice relaxation times and 48 Table 2. Pulse methods for relaxation studies. Methoda Pulse3sequenceb Reference IRFT(T1) (180x-mT-90x-FID-ST1)n 82 SRFT(T1) (QOx-mT(FGP)-90x-FID)n 83 PSFT(T1) (QOX-FID-ml’)n 84 FIRFT(T1) (180x-mr-90x-FID-T)n, T<5T1 85 SSFT(T1) 180x—(I-6-FID)m 86 Rapid PSFT(T1) (em-FID-T)n 88 cp (T2) 90x-(T-180x-T)m-FID 120 CPMGCTZ) QOX-(T-18Oy-T)m—FID 121 FFS(T1) QOX-T-90_x-T 590 Triplet (T1,T2) 180x-(t-90x-1-180x-T-90x)m 122 180x-(t-90x-1-180_x-T-90x)m 123 lsny'(t'gox‘T'180y'T‘90—x)m 123 Quartet (T1,T2) 180y-(t-QOX-I-180y-21-180y-I-90-x)m 92 Double Triplet 180y-(t-(90x-1-180y—t-90_x)2)m 92 aSee text for the meanings of the abbreviations. bm Denotes the number of observations; n denotes the number of scans for a signal accumulation; T denotes the variable delay; x and y are used to specify the direction of the rf pulse applied. 49 found that in weakly-coupled spin systems there might be an error of ~10% induced in the T1 measurement and that this should be more serious for strongly-coupled systems. Vold and Chan99 have studied transverse cross—relaxation which sometimes causes T2 to be longer than T1. A general study of transverse cross-relaxation in heteronuclear coupled AXn and AXY systems has also been reported by Vold and Voldlos. Recently a large amount of work on the theories of cross-relaxation and cross— correlation effects in 13C NMR relaxation studies has been reported by Grant and his coworkersllS-llg. To see these effects, let us consider a simplest two-spin case. Using Redfield's master equation, Equation (48), we can write where Sii(t)=Tr{Iz_°;} and mi is the Larmor frequency of spin i. The 1 Riijj's are the elements of the relaxation matrix and are represented by the following equations: - l. . R1111 ' 3J12(w1‘“2)+J12(w1)+2J12(“1+w2)+R1 (94) R --1-J ( )+J ( )+2J ( + )+R' (95) 2222 ' 3 12 “1'“2 12 “1 12 “1 “2 2 R - R - la (w w )+2J (w +w ) (96) 1122 2211 3 12 1 2 12 1 2 ’ where Ri and R5 represent the nondipolar random field contribution to thezrelaxation of each spin. Equation(93) can be rewritten as 50 s (t) iw +R R S (t) [.11 ] g _ [ 1 1111 1122] [ 11 ‘] (97) 522(1) R 1w2+R 822(t) , 2211 2222 The solution of Equation (97) can be obtained in case the initial conditions are known. If spin 1 is observed by applying the inversion- recovery method, then (denoting 11 for 1111, 12 for 1122 and 22 for 2222) 31.5? = -28?(b1exp(-a+t)+(l-b1)exp(-a_t)) (93) 82-53 = 28$b2(exp(-a_t)-exp(-a+t)), (99) o . . . . . where the 81's are the magnet1zat1ons at thermal equ111br1um and _ 1 1 2 2 ‘1 at ' 2{R11+R22)i(Z{R11’R22) +R12) (10°) b1 = (a+-R22)/(a+-a_) { (101) b2 = R12/(a+-a_). (102) It is obvious that each spin will have its magnetization decay bi- exponentially unless the off-diagonal element of the relaxation matrix, R1122 is zero. In two-spin systems the cross-correlation is absent, but in three-or four-spin systems there will be spectral density functions involving three spins included in the elements of the relaxation matrix117 EXPERIMENTAL I. The NMR Spectrometer A. General Description of the Pulse System The NMR spectrometer used in this study is composed of a Varian DA-60 spectrometer system with 12 inch magnet, a regulated power suply to generate a 14.1 kG field and a Varian V-K3506 Super Stablizer. To this has been added a highly modified NMR Specialties MP-1000 pulse system including two P-118A rf units, a P-103A power amplifier, a P-lOOA receiver and P-llSA, P-114A power supply units, a homebuilt wideband power amplifier, a few homebuilt tuned and wideband probes and a Nicolet 1083 minicomputer with several homebuilt interfaces. Observation of 19F nuclei utilizes all tuned components operating at 56.45 MHz. The observation of other nuclei utilizes a multinuclear system modified as described by Traficante gt_§13125 For long-time signal averaging, a field/frequency external lock system is used. The block diagrams of the overall system for 19F observation and multi- nuclear observation are shown in Figures 2 and 3. The detailed des- cription of the multinuclear system will be given in the next section. For 19F observation, the rf source for both the transmitter and the reference of the phase detector is a frequency synthesizer. The rf is sent to the rf unit in which it is amplified by two buffer amplifiers composed of two 7587 Nuvistor tetrodes in a push-pull configuration. The outputs from the rf unit are gated by two external. dc pulses ( 0 V to ~15 V) then sent to the P-103A power amplifier to $1 52 .cowum>uomno mafi pom gaumxm omfism ecu mo awhuuwc mooflm .N ouauwm >._aDm mam EH" Vx< Vx‘ - v2 om: .. mm . ~12 on m 7 82 wom :2: “a 1 on”: 53 .aoumxm amaze nmofluacwuasa may mo aauwmwc xoon .m enamHm HOOV.Zm wthh< . . . 5:: u w. _ .5 0%. no.2. " s “.— ~z2 on _ _ - _ . h u _ .hwnu “ " mm<1a , — q .- “Thzhil " . u- out I n 82 u _ $50.2 3.3.30.— . n , N128 " n : w W 5:92. _ 35.55% . - _ u _ .4‘ .m _ _ . . . _ . . _ _ \(W _ a 2 i S4 reach to at least 100 W. Only the low power stage which contains a 3E29 tube is used for this work and it is also gated by the same dc pulses. The arrangement of two crossed-diode boxes and a few quarter wavelength cables is to isolate the receiver from the transmitter so that the preamplifier would not be saturated or damaged by the strong rf pulses while transmitting. The signal induced from a single-coil probe is sent to a tuned preamplifier then to the P-lOOA receiver and then is phase detected. The detected dc signal can be further ampli- fied or directly sent to the computer in which it is digitized and stored. B. Multinuclear Modification The modification of the tuned pulse system for observing various other nuclei is by means of a simple rf mixing technique. Figure 3 shows the diagram of this multinuclear system. The rf unit P-118A is equipped with a crystal oscillator which provides the 19F Larmor frequency 56.4 MHz. The frequency synthesizer is used to provide a mixing frequency which is the difference between the observing fre- quency and 56.4 MHz. The attenuator after the rf unit is used to match its output amplitude with that of the mixing frequency. The mixed rf is sent to two power amplifiers, an ENI-406L and a homebuilt PA. The final output is at least 100 W which is sent to a wideband probe with a single-coil configuration. The signal picked up is carried by the observing frequency which is further mixed back to 56.4 MHz so that it can be detected by the tuned receiving components. All the rf mixings are carried out by a few double balanced mixers 55 ZAD-l (from Scientific Components Corp.). Two wideband preamplifiers are used before the receiver and an image filter is used to filter out any higher harmonics. C. The Probes The advantages and disadvantages of using a crossed-coil or single-coil-probe have been discussed by several authors73’74’126’130. The advantages of the crossed-coil probe are the minimum rf leakage and the optimum sensitivity attainable, but the power efficiency is less and the mechanical complexity is greater than a single-coil probe. The latter has, however,a serious rf leakage problem which must be overcome by more delicate circuit designlzg. In this work, the probes used are . all single-coil type. For 19F observation, the probe is made for 5 mm sample tubes. It has a coil wound on a piece of 8" long, 0.25" i.d. quartz tube with seven turns of #26 magnet wire and tuned to 56.4 MHz 131 with Q approximately equal to 30 . In order to do a few multiple— pulse experiments which need a field-gradient pulse, for example, T1 132 measurement using the homospoil technique , modified hybrid relaxation time measurementlss, self-diffusion coefficient measurement using the PUISed field-gradient spin-echo techniquelM'147 , etc., two types of field-gradient coils were constructed in the probe. The first one is a pair of anti-Helmholtz coils with coil axes pointing against each other along the direction of static magnetic field. Each coil consists of 15 turns of #30 magnet wire wound on a Delrin form with the configu- ration proposed by Tanner135 to produce the most linear gradient at the 144-146 sample (see Figure 4a). The second one is a quadrupole coil built on a cylinder (see Figure 4b) which can slide over the sample 56 \ .< Figure 4. Field-gradient coils in the probe: (a) anti-Helmholtz configuration; (b) quadrupole configuration. S7 coil. Each quadrant consists of S strands of #30 magnet wire which can easily be increased to provide a higher field gradient. The latter type has several advantages over the former in that it is relatively easier to construct, the configuration is not so critical to the linearity of the field gradient produced, the main component of the field gradient can be changed to a different direction simply by rotating the cylinder and the dc resistance and the inductance of the coil are less so the eddy current formation is minimized and a higher field gradient is attainable for equal current supplied144’145. The wideband probe for observing other nuclei is of the insert- type. The coil is wound on a cylindrical quartz tube which can slide into the probe body in which the other part of the probe circuit is permanently built. In order to keep the impedance matched through a wide range of the observing frequencies, 2 MHz to 25 MHz, the coil is tapped at about one turn from the end. The exact tap point is deter- mined by several tries until the impedance is measured to be 50 ohms. 13C, 14N and 20 The detailed circuit diagram is given in Figure 5. For observation in this work, the insert was made by winding 24 turns of #28 magnet wire on a 0.5" o.d. thin-wall quartz tube to give ~12 mm coil length. The tap point was made at one turn from the top and the impedance was measured to be from 30 to 75 ohms for frequencies from 4 to 20 MHz. Care has been exercised to obtain a uniform coil spacing and pitch. With this insert, a capacitor pad (220 pF) is needed for14N observation. The Q of the circuit is found to be 30 to 50 depending on the frequency operated. With about 300 W transmitting rf the 900 pulse width is about 8 usec for 13C, 15 usec for 2D and 30 usec forl4N. 58 ? TO PREAMP '4? : llll lllll T f he Figure 5. The inside cross-section of the wideband probe: a, sample tube; b, spinner; c, BNC connector for the capacitor pad; d, H1 tuning capacitor; e, heater sensor connector; f, N2 gas inlet; g, H1 rf cable; h, heater sensor; i, copper- constantan thermocouple; j, decoupling coil; k, sample- coil; 1, external lock probe; m, decoupler tuning capaci- tor; n, LC circuit of the probe. 59 D. Computer Interfaces In order to have the computer drive the spectrometer, an interface must be used to convert the logic levels of the triggering pulses generated from the computer to those required by the spectrometer. Also, since in this work the computer can only provide triggering pulses with accurate time intervals but not the duration of the pulses themselves, the interface should also have the function of adjusting the pulsewidths. A well-constructed interface131 is used for these purposes. This interface converts two I/O (input/output)pulses from the Nicolet 1083 computer (0-3 V, 500 nsec) to two 0-15 V pulses with pulsewidth adjustable from 1 to 100 usec. In addition, a T1/T1p switch control can convert the second pulse to be a terminating trigger for a long pulse activated right after the first pulse. A detailed circuit description was given by Wright131. In those experiments which needed field-gradient pulses, a cons- tant current supply was used. Figure 6 shows the circuit diagram of this unit. It contains a pulsewidth adjusting circuit based on a mono- stable multivibrator SN 74121; the adjustable range is from 100 usec to 10 msec. The output from pin 1 of the SN74121 is used to switch the transistor 2N3415 off. A voltage divider is used to control the base voltage of the transistor 2N3053, which in turn controls the current passed through a power transistor 2N6OS7. The maximum current attain- able is about 10 A. In hybrid relaxation studies, the pulse sequences require the adjacent 900 pulses to be 1800 out-of-phase with each other. This is achieved by use of the phase shifter circuit shown in Figure 7. The phase shifter is triggered by the same pulse, PULSEZ, as triggers 60 .m m: a“ co>ww one mnoufioamao .mano cw co>wu one muoumwmonv memsm unounau ucowweHm-pHowm awash on» we aahuewv ofiuaaocom .o onsufim stoo >n—— "V'V' our 61 n.0mwzuonuo penance mmoacs .mn :« one muouweemmo Ham .mazo cm can muoumwmon afimv «macawo mcfimno>ou omega my on» we seaweed owuaaunom .n ousufim v: 4&6 . AAAAAA V'V'V‘ l. T .50 “E A 109 fill. noonz N 51000” , Q o o — .. 1 -§ § 2 1 Zoo onn Own v.0— noomz m/r ....:.: _ 2.2. I... Qoo .: v.3 "U'V' 4H:- 1.2+ 62 the 900 pulse. The output from the flip-flop SN74109 is converted into :15 V and acts as the gating pulse of a double balanced mixer which all the rf pulses pass through before going to the power amplifiers. E. External Lock System When performing a long-time signal averaging, the magnet field has to be very stable. This is achieved by a homebuilt external lock system which uses the original DA-6O console as the lock circuit. The lock probe was made of a single coil tuned at the proton frequency131. The lock sample used is water doped with copper chloride. A more detailed description has been given by Wright13l. F. Temperature Control and Measurement The temperature of the probe is controlled by a Varian V-4343 temperature controller and a Varian heater-sensor inserted into the Dewar. A stream of dry N2 gas is precooled or preheated before being sent to the probe. Temperature detection is achieved by using a copper-constantan thermocouple with the junction lead placed about 0.5 cm below the bottom edge of the coil (see Figure 5). The actual temperature of the sample is calibrated relative to the apparent readout by putting another thermocouple directly into the sample tube. The measured difference was found to be less than 0.50K. The accuracy of the temperature measurement was within i0.5°K except at very low temperature ( below -100°C) where it was iloK. 63 G. Heteronuclear Decoupling and the Variable Field Lock System Heteronuclear decoupling in 13C NMR provides many important 148'151. In relaxation studies, the cross-relaxation effect advantages due to dipolar interaction between nuclei can be eliminated by the 113 decoupling technique . In addition, the nuclear Overhauser effect (NOE) can be measured148’149 and the dipolar relaxation rate can be estimated. The requirements for heteronuclear decoupling are normally a decoupling coil built over the sample coil and a decoupling rf source. In this work, the decoupling coil is a Helmholtz coil wound on a 0.75" o.d. quartz tube which is slid over the observing coil. In order to do a dynamic Overhauser effect experiment152 , a decoupling box is made which provides a band of closely-spaced pulses to trigger the de- coupling rf. The purpose of using closely-spaced pulses rather than a cw pulse is to minimize the duty cycle of the rf power amplifier P103A. Figure 8a shows the schematic of this decoupling box. The dual mono- stable multivibrator SN74123 acts as a self-triggering pulser which is turned on and off by the output of flip-flop SN74107 which is in turn controlled by two input pulses. The final output.pulses are all ampli- fied by common-collector transistor amplifiers which have a voltage gain of approximately unity but a large current gain. The timing diagram is also shown in Figure 8b. The modified setup for the heteronuclear decoupling experiment is shown in Figure 9. An additional frequency synthesizer is needed for the decoupling rf source. For high power decoupling, better rf isolation can be achieved by the time-shared mode provided by the decoupling box which blanks the receiver while the decoupling rf is 64 4 5 1;?5 +5 —15 4 T21: 1000 ‘1! . )4 )1 2Q5 20K 'K 1471,25K Cm o——-§1 220011000 K I ' o 4 74121 16 13 3'4] 7 '01 75K 75" 2N3¢45 7 1? 4 ' 2N2222 10 - ‘ +5 7 4123 ' m 3 7 8 ? - Bin 9 14 n 9 +5 130 1K 74107 a . F s PULSE B A C B A C Pulse . Sequenccl T k 1: . T K I I’ 01.22:: GNU H H H H ,1; II V 08': —1' flflfl H ...................... r Figure 8. (a) Circuit diagram of the decoupler gating box (all' resistors are in ohms, all capacitors are in uF). (b) Timing diagram of the outputs. 6S PULSE A Pulse width to normal trans- controller mission line to receiver PULSE B V “"c Decoupler P Receiver PULSE C Gating Box r Gate RI" I PA Gate Cato signal Decoupler L R to decoupler \ 1 ———o RF Source DBJ P°A' coil Figure 9. Block diagram of the heteronuclear decoupling setup. V 4311 RF UNIT ' [PR5 Mi Figure 10. 66 l l _ “2‘21 1 TIME SHARE BOX MOD Block diagram of the variable field lock system. 67 transmitting. It has been demonstratedlss’154 that by varying the frequency information about the contribution from the chemical shift anisotropy and from the scalar coupling to the total relaxation rate can be obtained. In order to make the varied field stable enough to do long- time signal averaging, a field/frequency lock system was also construct- ed. It consists of a wideband probe with a tapped coil tunable from S to 30 MHz (aided by an external capacitor pad), a lock sample contain- ing HZO/DZO (l;4 ratio by volume) doped with nickel chloride and a frequency mixing network similar to that used in multinuclear observa- tion. Figure 10 shows the diagram of the variable field lock setup. . 11. Measurement of the Relaxation Times A. NMRLX-- A General Program for Computer Controlled Pulsed NMR Relaxation Studies All the relaxation studies in this work were performed by using a general program NMRLX written for a Nicolet 1080 series minicomputer. The detailed instructions for using this program are given in Appendix A. For linewidth and second moment measurements, program SECMFT was used. The instructions are given in Appendix B. Several authors have studied the systematic errors in NMR relaxa- tion studies caused by the static magnetic field inhomogeneitylss, the H1 inhomogeneity156-159, the rf phase error160 and the rf pulse 161 defects . Since NMRLX provides a suitable way to determine the rf pulsewidths and the phases with well-adjusted field homogeneity, the 68 measured relaxation rates in this work are accurate to :38 for T1 values and 15% for T2 values except for some 13C Tl's which were only measured with ~10% confidence due to the small S/N ratio (~20) obtained. The reproducibility was found to be normally within 5%. 8. Sample Preparation All the samples were commercially made except CDBrS which was prepared previously131. The samples were all purified by one or two vacuum distillations (or sublimations in case it was solid at room temperature) after being carefully degassed with several freeze- pump-thaw cycles. After purification, the sample was distilled or sublimed into the sample tube until the proper amount was received then the tube was flame sealed. The sample tubes were specially designed using standard-wall Pyrex tubing. Since most of the compounds were studied from the solid phase to about the critical point, the pressure inside the tube varied from 0 to ~40 bars. In order to avoid an explosion while making a measurement, each sealed tube was tested at 10 to 20° above the highest operating temperature in a shielded hood. Figure 11 shows the design of sample tubes. The observing chamber was made to be spherical in shape and the size was no bigger than the coil dimension so that all of the sample compound was inside the coil. Above the observing chamber, a narrow pathway was also incorporated to prevent extensive convection and diffusion to the gas-liquid.interface. For some com- pounds, a second seal-off was necessary to avoid sample rapidly distilling from the bottom to the top of the tube and so causing bumping and large thermal currents. 1 N2 in 41 Figure 11. Design of the NMR sample tube. RESULTS AND DISCUSSION 1. Melecular Motion and Relaxation Studies of Some Small Molecules in the Liquid Phase A. Anisotropic Reorientation and 13C, 14N,19F Relaxation Studies in CF3§N_ 1. Introduction Due to the complexity of theories describing molecular motions in dense liquid phases,experimental evidence has usually been obtained from studies of simple and highly symmetric molecules which can largely simplify the theoretical derivations and also decrease the errors due to the approximations usually needed for larger and more unsymmetrical molecules. Many of the NMR relaxation studies have been done on spherical- top, linear and symmetric-top molecules for these reasons. For spherical and linear molecules, molecular motions in the liquid phase are mostly isotropic. However, in symmetric-top molecules the motional anisotropy might be large for those with a strong dipole moment and highly aniso- 162,163 164 and CHSI . In these molecules, two parameters are needed to describe the reorientational tropic molecular shape, e.g., CHSCN motion, therefore at least two nuclei at different locations should be studied68. Since, for spin 8 nuclei there might be more than one contri- bution to the total relaxation rate, the best choices will be the quadrupolar nuclei in which there is only one dominant relaxation mechanism to be considered. Unlike CH3CN,and its chlorine derivative CC13CN, trifluoroacetonitrile CF3CN does not possess a second quadru- polar nucleus. Therefore the elucidation of reorientational motions in 70. 71 CF CN depends on success in separating the different relaxation 3 mechanisms of the 19F nucleus. We report here our attempts to dothis in Sections 2 and 3.The separated spin-rotational relaxation rate was then used to test the validity of extended diffusion mode1522-24, which are discussed in Section 4. A limited 13C relaxation study was carried out and is reported in Section 5. Finally, the relation between 13C-19 F . 13 19 . . . . coupling constants and the C or F chemical shifts of some Similar compounds and their derivatives is discussed in the last Section. 4 2. 1 N Relaxation and the TUmbling Motions The measured 14N spin-lattice relaxation rates in liquid CF3CN are listed in Table 3. These values are also plotted against 1/T in Figure 12. From the linear dependence of R1,N on 1/T the correlation times, and therefore the rotational diffusion coefficients, of the C3 symmetry axis can be obtained by using Equation (67). For 14N in CF3CN, this can be simplified to Q _ 2 . R1,N _ 1.233(Q.C.C.) T6 , (103) since the asymmetry of the electric field gradient around triplly-bonded 14N is zero. Q.C.C. stands for the quadrupole coupling constant equ/h. From a microwave study of gaseous CF3CN, the Q.C.C. of 14N was found to be -4.70 MHzl66. Several authorsl62’167'169 of 14N, 17O and 2 have argued that the Q.C.C.'s D in liquids are usually very close to the values found in the solid phase but are smaller than those in the gaseous phase by m6 to 14%169. Therefore, a value of -4.2 MHz (10% lower than -4.70 MHz) was used for liquid CF3CN. It should be noted that using other values might Table 3. 14 72 N spin-lattice relaxation rates in CFSCN. T lOOO/T R1 T lOOO/T R1 1°C) (°x'1) (sec 1) 1°C) (°x'1 (sec-1) 12.0 75.51 222111 -33.0 4.17 404:8 -1.0 3.68 276:13 -34.0 4.18 396:13 -2.0 3.69 277:4 -44.0 4.37 452:12 -18.0 3.92 314:4 -88.0 5.40 908:34 -20.0 3.95 315:7 26.0 3.34 198:10 -32.0 4.15 350:8 -92.0 5.52 904:70 -47.0 4.42 449:8 -93.0 5.56 1021:81 -48.0 4.44 461:7 -103.0 5.88 1172:55 -61.0 4.72 543:8 -ll8.0 6.45 2089:90 -62.0 4.74 547:6 -24.0 4.02 337:10 -65.0 4.81 567:7 -3.0 3.70 267:7 -75.0 5.05 675:11 3.0 3.62 219:13 -88.0 5.40 834:17 8.0 3.56 196:6 -86.0 5.35 779:14 -5.0 3.73 225:20 -106.0 5.99 1228:34 ~123.0 6.67 1818:101 -115.0 6.33 1539:52 -107.0 6.02 1023:98 -123.0 6.67 2667:70 -93.0 5.56 746:14 -129.0 6.94 3136:330 -84.0 5.29 630:30 -124.5 6.73 2286:90 -52.0 4.52 379:37 -124.0 6.71 2159:123 -73.0 5.00 662:30 -116.0 6.37 1614:31 -61.0 4.72 538:19 -114.0 6.29 1580:61 -45.0 4.38 446:17 -100.0 5.78 1046:33 -20.0 3.95 371:13 -99.0 5.75 1080:50 -6.0 3.74 310:11 -90.0 5.46 890:19 25.0 3.36 210:10 26.0 3.34 208:12 -0.5 3.67 254:11 12.0 3.51 244:10 -78.0 5.13 682:32 -6.0 3.74 320:10 -94.0 5.59 1207:97 -20.0 3.95 355:8 -107.0 6.02 1622:144 -21.0 3.97 347:14 -58.0 4.65 533:26 73 201$ 10 R1 (s e c") § _ h 1 I l 4 5 6 7 8 1000/T Figure 12. Plot of the 14N spin-lattice relaxation rate in CF3CN as a function of reciprocal temperature. 74 change the re '5 calculated, but the activation energy AEL would not be 1 affected. The calculated 1 's and D 's ( D = ———-) are lotted in . 61- _L ‘L 6T6 P Figure 13 and the activation energy is found to be 1.40 kcal/mole. 3. 19F Relaxation and Anisotr0pic Reorientation l9 . . . . The measured F spin-lattice relaxation rates are listed in Table 4 and plotted against reciprocal temperature in Figure 14. It is . . . 1 . eVident that several mechanisms must contribute to 9F relaxation, and we may write d,intra+Rsr+Rcsa+Rsc t _ d,inter R'R * 111’ 1 1 R (104) where the first and the second terms are the intermolecular and intra- molecular dipolar relaxation rates, respectively, the third term is the spin-rotational relaxation rate, the fourth term is the relaxation rate due to the chemical shielding anisotrOpy of 19F nuclei and the last term is the 19F scalar relaxation rate due to» the 14N nucleus. From Equation (69), the contribution from the scalar coupling JFN can be estimated as sc _ ~16 2 R1 - 4.QXl0 JFNR1,N' (105) Very few J values have been reported but, since R is about the FN 1,N order of 103nsec'l, Ric is negligibly small and can be completely ignored. csa R1 can be estimated from Equation (73a) which is, in this case, 7S 50- 79.1. DI 1 ‘1 o. "'5 F5 701 A 9. l a V g? 1 .— 0.5— 1 4 J L 1 1 4 5 6 7 8 1000/1' Figure 13. Reorientational correlation times and diffusion coeffi- cients of CF3CN as functions of reciprocal temperature. Table 4. 19 76 F Spin-lattice relaxation rates in CFSCN. T 1000/T Rl T 1000/T R1 1°C) (°x‘1) (sec' 1 (°C) (°K'1) (sec 1) 25.0 3.36 0.724: .010 -116.0 6.37 0.208: .002 21.0 3.40 0.709: .003 -120.0 6.54 0.225: .003 12.0 3.51 0.621: .009 ~126.0 6.80 0.260: .002 0.0 3.66 0.517: .005 -131.0 7.04 0.292: .006 -11.0 3.82 0.455: .008 -131.5 7.07 0.309: .006 -20.0 3.95 0.412: .007 -135.5 7.27 0.373: .011 -29.0 4.10 0.371: .006 -l41.0 7.58 0.452: .007 -38.0 4.26 0.322: .002 -140.5 7.55 0.438: .006 -37.5 4.25 0.323: .003 -145.5 7.84 0.572: .004 -52.0 4.52 0.271: .011 -148.0 8.00 0.698: .007 -53.0 4.54 0.268: .004 -151.0 8.20 0.896: .020 -57.0 4.63 0.258: .003 -153.0 8.33 1.027: .020 -62.0 4.74 0.247: .005 -154.5 8.44 1.172: .009 -66.0 4.83 0.241: .003 -156.0 8.55 1.301: .009 -68.5 4.89 0.237: .002 -157.0 8.62 1.452: .012 -73.5 5.01 0.230: .003 -158.0 8.70 1.550: .020 -77.0 5.10 0.213: .002 -158.5 8.73 1.630: .013 -77.5 5.12 0.205: .002 -159.0 8.77 1.630: .008 -82.0 5.24 0.199: .004 -88.0 5.40 0.192: .003 -92.0 5.52 0.185: .002 -94.0 5.59 0.190: .002 -96.0 5.65 0.189: .001 -99.0 5.75 0.189: .001 -103.0 5.88 0.190: .001 -lO7.0 6.02 0.193: .002 -1ll.0 6.17 0.200: .001 0.5 *- nbemv .m m. 0.05 ~ Pi 77 0.05 - l 1 l 6 7 8 4 5 1000/1 Figure 14. Plot of the 19F spin-lattice relaxation rate in CF3CN ' as a function of reciprocal temperature. 1. For cont buti Thel 74% 1 whet EqUa tOt ”he (0. 78 4 2 RC53 = 1.68x10 A0 1 . (106) 1 ppm9 For Aoppi~200 ppm, Risa would become ~58 of the intramolecular dipolar contribution. Since A0 is not exactly known, we will ignore this contri- bution for the time being and discuss it later. The intermolecular dipolar relaxation rate can be estimated from the densities and the self-diffusion coefficients as summarized in Equations (59), (60), (62) and (63). Gillen17o has discussed the use of different numerical factors and proposed that the best one is d,inter _ wv‘fizN R 1 4Dsa (107) The molecular radius a can be estimated from the density by assuming 74% of space is occupied by molecules and so a = (0.74 M) 1/3 (108) 4WNod where d is the liquid density and No is Avogadro's number. The N in Equation (107) is the number of spins per unit volume and is also related to the density d by Nodno g ________ , 109 Mol.Wt. ( ) where no is the number of spins per molecule. In order to take the pressure effect on the density into account (the pressure in the sample tubes can vary from ~0 bar at the melting point to ~40 bars at the highest temperature studied which is ~10°C belo‘ dire: weig! in T: visc< may c There spin- Equat plot! rates dipol their deper have the 1 low ( P10t1 Corre Equa- wheT 79 below the critical point of CF3CN), we have measured the liquid density directly in the sample tube by carefully calibrating its volume and the weight of the sample sealed in it; these density measurements are listed in Table 5. The self-diffusion coefficient can be estimated from the shear viscosity of the liquid sample but it has been found that Ds so obtained may deviate from the direct experimental result by 20% or more171 Therefore, we have directly measured Ds by the pulsed field-gradient 13 spin-echo technique 8 and these values are listed in Table 6. Using d,inter 1 plotted in Figure 14. By subtracting R Equations (107), (108) and (109), R values were calculated and are d,inter 1 rates, we can obtain curves representing the sums of intramolecular from the total relaxation dipolar and spin-rotational relaxation rates. d,intra and Rsr their opposite temperature dependence. Since the low temperature linear The separation of R can be achieved by making use of dependence of relaxation rates is not available for this compound, we * have assumed that the reduced reorientational correlation time 16 at d,intra so that the 1 r form a straight line; this best fit is the critical temperature is ~0.7 and have adjusted R 5 low temperature values of R1 plotted in Figure 14. The intramolecular dipolar relaxation rates are related to the correlation times of the motion of 19F-lgF internuclear vectors by Equation (58) which can be written, for this particular compound, as d,intra 10 R1 = 1.31X10 Te , (110) where T6 is related to the diffusion constants by Ta 80 Table 5. Measurements of the density of CFSCN. c d T v, v: 93 W8 w, d, (OC) (m1) (m1) (torr) (g) (g) (g cm'3) 18.0 0.755 0.6011 28,840 0.0913 0.6763 0.8958 -14.0 0.670 0.6861 10,965 0.0444 0.7232 1.079 -9.0 0.685 0.6711 12,882 0.0500 0.7176 1.048 -28.0 0.654 0.7021 6,607 0.0289 0.7387 1.130 -25.0 0.660 0.6961 7,413 0.0317 0.7359 1.115 —21.0 0.665 0.6911 8,511 0.0356 0.7321 1.101 -14.0 0.670 0.6861 10,965 0.0444 0.7232 1.079 -6.5 0.692 0.6641 14,454 0.0550 0.7126 1.030 3.0 0.710 0.6461 19,055 0.0682 0.6994 0.985 5.0 0.720 0.6361 19,953 0.0698 0.6978 0.969 8.0 0.727 0.6291 21,878 0.0748 0.6928 0.953 11.0 0.735 0.6211 23,988 0.0800 0.6876 0.936 14.0 0.747 0.6091 26,303 0.0851 0.6825 0.914 -60.0 0.600 0.7561 1,622 0.0088 0.7588 1.265 -53.3 0.605 0.7511 2,291 0.0119 0.7557 1.249 -48.0 0.617 0.7391 2,884 0.0144 0.7532 1.221 -46.0 0.622 0.7341 3,162 0.0156 0.7520 1.209 -43.0 0.627 0.7291 3,548 0.0172 0.7504 1.197 -41.0 0.632 0.7241 3,890 0.0185 0.7491 1.185 -75.0 0.571 0.7851 780 0.0043 0.7633 1.337 -160.0 0.490 0.8661 0.112 0.0000 0.7676 1.566 aVg = vtotal - Vi, where vtotal = 1.3561 ml (volume of the tube). blog Pv(torr) = 7.900 - lOOO/T, from Reference 302. cW = 95 P V 62.3T. 8 v g/ d = - W, Wt Wg , where Wt= 0.7676 g.(total weight of the sample). Table 81 Table 6. Self-diffusion coefficients of CFSCN. a b T 1000/T g 05d 0 o ‘1 -5 2 ( C) ( K ) (G [Cm) (10 cm /sec) 17.0 3.45 150 4.37 5.0 3.60 158 3.61 -6.0 3.74 170 2.98 -21.0 3.97 185 2.44 -37.0 4.24 208 1.59 -51.0 4.50 232 1.29 -71.0 4.95 282 0.800 -73.0 5.00 288 0.785 -87.0 5.38 337 0.519 -103.0 5.88 418 0.310 -104.0 5.92 422 0.294 -118.0 6.45 531 0.171 -131.0 7.04 684 0.0868 -136.0 7.30 760 0.0713 8Magnetic field gradients calibrated with CFClS. 13See Appendix A for the method of calculating Dsd’ 82 = (360529-1)2 + 35in26cosze + 3sin40 4 + 2401- SleDll (201-49'!) T0 (111) . . . . . l9 . . which is essentially the same as Equation (87). For the F nuclei in CFSCN 0:900, therefore 1 3 T0 ‘ 2710f 4(20r40ll) ’ Tei 'f(r), (112) where f(r) is a function of r, the anisotropy of the reorientational motion defined by r = DI I/DL = Te-L/Te . (113) By using Di or 161 obtained from 14N relaxation studies, we obtained D11 and T6 which are shown in Table 7 and are plotted against the reciprocal temperature in Figure 13. The activation energy for the spinning motion (reorientation about the C3 symmetry axis) was found to be ~0.89 kcal/mole which is smaller than AE1f~1.40 kcal/mole. This implies that the reorientation about the symmetry axis is easier than that about an axis perpendicular to the symmetry axis. From Table 6 one can also see that the anisotropy of the rotational motion is temperature dependent, varying from ~6 at -148°C to ~l.8 at room temperature. Similar phenomena have also been found in other symmetric-top molecules Sl63-165,172. which have the dipole moment along the symmetry axi It is of interest to compare the above result for CF CN with those 3 for CHSCN and CC13CN165. Table 8 shows some pertinent physical constants and motional parameters for these compounds. CHSCN has been widely 162,163,173 174-176 studied be NMR , Raman spectroscopy and dielectric 83 mm.m v~v.o m.mH 0v.» wo.m mw.o oHH.o nNo.o oo.w mNH oo.m ome.o o.v~ on.o oo.o om.“ ommo.o Nmo.o om.h mmH mm.v mwv.o om.m mn.v on.“ on." omoo.o oeo.o oo.n mefi nm.m wom.o mm.o em.n mN.m ov.m Novo.o moc.o om.o emu He.m wmm.o om.e mo.~ n.HH ~e.m oemo.o ~mo.o oo.o nos cm.~ mum.o mv.m mm.~ ~.v~ om.v oomo.o mNH.o om.m NwH mo.~ oHo.o mv.~ me.H o.w~ mm.c vmfio.o ~m~.o oo.m com om.~ mmo.o mo.~ mm.” m.o~ nH.m wo~o.o c-.o mu.e cam wn.m Heo.o -nn.u ~H.~ c.m~ mo.m oeHc.o ooN.o om.e NNN om.~ noo.o ve.m Ncm.o m.mN o.- o-o.o Nam.o mm.e mmm mo.~ vae.o ~N.H oem.o w.n~ h.nn oHHo.o vwm.o oo.e 0mm em.“ eon.o we." m~n.o m.Hm n.0H emoo.o mwv.o on.m com mm.~ onn.o om.o o~o.o H.0n m.m~ Nwoo.o m~0.o om.n own ~m.~ mnn.o om.o mam.o w.sm w.o~ whoo.o mon.o ov.n emu we.” ~e~.o 55.0 ~um.o e.wm o.HN mwoo.o av5.o mn.n mam eu.H Nmn.o en.o 5mm.o o.mm N.- mneo.o omn.o om.n non mflommv aoommv nuuoommauV “H-00mmaov annuomv nfiuoemv Amuuov axov o o a have 0 P oo~-o~ acH-oH unpem.wm awe e\ooo~ e .zommu you mnouoaepmm demeanoe camouuomfic< .n oanmb .r‘h uri‘w.ut;\tnwrcd.lql~|mdmmuvu$.ll OLH vcm 0~mhuwc0uoum mo wheuefimhmk HmCOMuOE 32¢ mucmumcou ~GUHWXCQ QEOW ho :OmmthEoU .m 0~nmb 84 .ANQmHV HNe «mm .memmom .amsu .oom .H~:m .eoflox use cm>n .mmomfiw omen .oom .aeau .H .owaouwm .o .a .0 map H90 .6 .m .eh>emoa .3 .6 .mm .uomm pa mosfim>. =N.Qfle.~ mH.m me.e mefi.aflwo.~ mm.qflm.H mw.oae m~.nfim mme.o uee.~ amm.H zomfiou ee.wm eo.HN eoe.~ cam.o cw.am~ gm.oeH --- emmm.o nmm.H zunmo afi.m m.ouo.~ :mm.~ xe.~Hw.mH xfl.ouw.fi hm.o“w.o mm.co~ www.cfi zonao xmwuamfi sm.oH oo.~ one.o mHH.Hm mev.m omvm.o umaa.o awn.» zommu Hnoon Nae fl.oHoa amux mmao my “mac my meow Amuse my new H_ao~o~ Manofiofi .Am __m AWoeoH __Hoeo~ .c .e a announce .efifiuuw=0pooeofim:wnu 0&9 awn—d QHMHHMCOHQUQ MO WHOHOEMHQQ HGCOfiHO—n awn—m mHGmHmCOU Haofimxgm oaom mo commuemaou .m manna 85 .xuoocu Anemoomfl>onowe 0;» Beam commence moafim>c .nnm oueonemoma .05” ooconemomu .mufi monogamomx .Noa ooeouomomn .Aammmv can“ 4mm .msna .Eocu .5 ..un .82“: .u .m can ”can .5 .Um .mmm oucouomox: .mommav em «mm .>om .maca .seaoo .2 van oflspansane .m .mcmm .: .nofimmae .zm .mcn Oucohmmmmm .mofi oozonomoxo .JHOI mwflFfl .AN .~o>VfimomHU "as .Ho>vnommfiv xuo» zuz .ue«>ommm .HH was H.uHo> .:mv::oQEou owcmmno when we mueeumeou Heowao:Unoofimxnm: .meeahoaafih .ho 86 relaxation measurement177. The dipole moment u=3.38 D . and the'an- isotropic molecular shape (ll/III~10),of CHSCN make its reorientational motion very strongly direction dependent. At room temperature, the spinning motion is about ten times faster than the tumbling motion. For CCISCN, the dipole moment of 1.99 D and a nearly spherical molecular shape make the motional anisotropy much smaller ( about 2 at room tem- perature). For CF 3 the anisotropy of the reorientational motion is about 1.78 at room CN the dipole moment is 1.33 D and Il/Ill~2.0 so temperature. It can be concluded that not only the molecular shape but also the dipole moment of a molecule determine the reorientational motion. Different activation energies for parallel and perpendicular reorienta- tion are also evident in these compounds. While the activation energies AEL for CHSCN and CC13CN are found to be in accord with those obtained from Raman and dielectric relaxation studies and the predicted values from microviscosity theory211, this comparison is not available for CF3CN since the other studies have not yet been made. The smaller 0E1- (1.40 kcal/mole) for CF3CN compared to that of the lighter analog CHSCN (AE1~2 kcal/mole) is probably due to the much smaller dipole moment of CF3CN. Since the dipolar "head-to-tail" attraction tends to align the molecules in a unique direction and interfere with the tumbling motion of the molecules, this effect should be much stronger in CH CN than in 3 CF3CN. An overwhelming contribution of the spin-rotation interaction to the relaxation rates is typical in small fluorocarbon;131’171’178'184. Since in CFSCN the 19F nuclei are not on the symmetry axis, the princi— pal coordinates which diagonalize the spin-rotation tensor will not diagonalize the diffusion tensor or the moment of inertia tensor 0f the 87 molecule. Therefore, Equation (82) should be used for relating Rfr to the motional parameters. Unfortunately, there are too many unknown variables in Equation (82) and it becomes almost impossible to use. Despite this, we will attempt to use a simplified equation (Equation (76)) to derive the overall angular momentum correlation times. Equation (76) can be written as 8n2 2 39 sr _ k _ 2 R - ( ”2 )TIavCefftJ _ 9.8x10 TIavCeffrJ , (114) 1 where Iav is the average moment of inertia and Ceff is the effective spin-rotation constant in Hz. The spin-rotation constant can be obtained . from the molecular beam experiments or can be estimated from the para- magnetic chemical shielding. Flygare185 and Deverall186 have derived an equation to relate the spin-rotation constant with the paramagnetic part 187 of the chemical shielding based on Ramsey's equation . A general form of this equation is as follows: 06 = Cabs - 0d = K-loca , (115) where cabs is the absolute chemical shielding of the observed nucleus, °d is the diamagnetic chemical shielding and K is a nuclear constant 30 19 which is 2.09X10 for F. 0d for 19F has been calculated by the atom- 88 dipole method1 and many different values have been obtained but the 1 most usually used seems to be 471 ppm 89. Thereforel71, 29 - 471 = 6.97 10 2C I C to... 1,... " ‘ 11* IIIII) 29 ’ W 6.97x10 3IavCo. (116) 88 19 The F chemical shift of CF CN was measured and found to be 59.7 ppm 3 relative to CFCl3 which has oab§~195.6 ppm188. Therefore, the spin- rotation constant C, can be calculated from Equation (116) and we obtain Co .-4,29 kHz. Using this value for Ceff’ TJ'S can be obtained and are listed in Table 9. Since rJ is the correlation time describing the motion of the angular momentum, which is changed each time a collision occurs, 1/1J can be considered as the collision frequency. On the other hand, the reorientational correlation time recan be considered as the time needed for a molecule to reorient one radian. It can be seen from Table 8 that at low temperature, e.g., 1540K (~119°C), a molecule will undergo a collision once for every 20 reorientation. Therefore, the diffusion limit or small-diffusion-step approximation is probably suitable for describing the molecular motion at this temperature. At higher tempera- tures, e.g., 2980K (25°C), a molecule has rotated 600 before it under- goes a collision. Therefore, a strong inertial effect should be consi- dered in describing the rotational motion of this molecule at higher temperatures. Another way to test the inertial effect is by comparing the experi- mentally derived reorientational correlation time with the theoretical reorientational correlation time of free gaseous molecules rf{~%/I;7ET . This test has been called the X test190 and has been widely used to test the applicability of the small-step diffusion model. x1 is defined by Te. - —-L w (117) X1 3 SJIi/RT where i stands for components x,y,z,or II and l_in this case. b _ . 89 eH.n omo.o mw.o ce.w Nvo.o oo.m mug me.m mvo.o nm.o om.o omo.o om.n mm“ o~.e noo.o om.o mn.e eno.o co.» neH ~m.m Hmo.o em.o mm.m nmo.o om.o em“ mm.m mNH.o wm.o mo.~ cmH.o oo.c 50H No.N eo~.c No.“ mm.H men.o om.m an mm.H e-.o no.“ ae.~ mom.o oo.m com He.H mo~.o OH.” wN.H He~.o mn.e oHN m~.~ vom.o Ma.“ o~H.H mom.o om.v NNN NH.H emm.o 0H.H ~0a.o mon.o m~.v mnN Ho.H eme.o o~.H oew.o nmm.o oo.v omm nww.o mHm.o mm." wan.o m~e.o mu.» com How.o oeo.o mm.” owe.o oom.o om.n emu vnn.o o~n.o om." mom.c Hmm.o oe.n «mm mvu.o cmn.o an.“ ~nm.¢ num.o mm.n mam mm5.o ”an.o ~n.n nmm.o mam.o om.n mom mfisuommv nuommv fluommv munxoy Agog m. m. £55»: 3 P .283 a. (r zunmu new woman cofiuemonnoo asucoaos Hmfismee pee Henofiueueofiuoon poozpox .m oflnmh 90 Figure 15 shows the plots of xi versus reciprocal temperature for both spinning and tumbling motions. In order to be in the diffusion limit, xi has to be much greater than 1. From Figure 15 we find that the re- orientation about the symmetry axis is about completely inertial except at very low temperature (~-l40°C); on the other hand, the tumbling motion of the molecules changes from the diffusion limit at temperatures below the normal boiling point (~-68°C) to the inertial limit at higher temperatures. Therefore, it can be concluded that the classical diffusion model can not be applied to this molecule except below the boiling point. . . . 165 Similar conclusions have also been drawn for several other compounds . 4. The Extended Diffusion Model For the purpose of bridging the classical diffusion limit and the free-rotor limit, Gordon22 has developed an extended diffusion model which has been discussed in the Theoretical part. Several tests have been made with small molecules191 and it was found that most of the molecules followed J-diffusion which is more reasonable in real liquids. In Figure 16, we plot the reduced overall reorientational correlation times 1; versus the reduced angular momentum correlation times 1;,which are obtained from the spin-rotational relaxation rates by using the spherical-top approximation. The result shows a line lying between the J-diffusion and M-diffusion curves at low temperatures and approaching J-diffusion when the temperature gets higher. Since we have assumed Ceff to be the same as CO, this might give a considerable error when the anisotropy in the spin-rotation constants is high. In fact, Ceff is always higher than Ca unless the anisotropy in C is zero. In order to 91 I l l 4 6 8 1000/T Figure 15. Plots of x , xi for liquid CF3CN. For comparison, the dashed lines are the corresponding x plots for CH3CN and the dotted lines are those for CC13CN. 92 make this correction, we considered the original derivation of Equation (115) by Flygarelss, who obtained the following equation for each com- ponent of the chemical shielding and the spin-rotation tensors, o = K-I -C 88' 8 88' ’ (118) where g=x, y or 2. Montana 35 al,192 have studied the 19F chemical shielding anisotropies of CF4, CF3C1, CF3Br, CF31 and CF3CC13 in smectic liquid crystal solutions and obtained the three principal components of the chemical shieldings in the molecular axis system. A very important conclusion they drew was that the substitution of an X atom for an F atom in CF4 had moved the principal axis of the chemical shielding tensor from the C-F bond to another axis by an amount which increased in the order ClHuoom -mon .mmou you :ofiuoonnoo «segue: one an“: mafiamon one moHono some new pfiaom ”Zummo you oawu newumfiounoo emueoaoa uofismem poospon msmno> oawu eofiuefiounoo Hoeowumpeowuoou pouncon mo uoHe .0" onswwm _ nd .0. «.0 .6 nod 36 _ 7 _ 1 _ _ 94 For the methyl carbon, only dipolar and spin-rotation interactions are important relaxation mechanisms. The dipolar relaxation rate R: due to three fluorine nuclei can be estimated as d__ 222-6 R1 - SYCYFH rCFTe , (119) where is the internuclear distance between C and F, which is rCF 1.33 3193, and 16 is described by Equation (111) with 0 for the C-F bond 69.3o 193; therefore, re~0.58 psec at 28°C and ~0.81 psec at -l4°C. From Equation (119) the dipolar relaxation rate of the methyl carbon 1 was found to be 0.0101 sec.1 at 28°C and 0.0140 sec" at -14°C. The spin-rotational relaxation rates are therefore equal to 0.024 sec"1 at 28°C and 0.016 sec‘1 at -1400. Since both 13C nuclei are on the symmetry axis of the molecule, Equation (81) can be used to relate Rir to the angular momentum correla- tion times. Equation (81) can be rewritten as Rsr - sflsz (I c2 T +21 02 r ) (120) 1 ’ "‘E?‘ J J ' an H" H H 1 where the spin-rotation constants CII and C1 are now in unit of Hz. In order to obtain Cll and Cl, another condition is needed. This can be obtained from the 13C chemical shielding by using Equation (115). For a symmetric-top molecule, Equation (115) can be written as . - .1 30 13 where K is 7.79xlO for C. We have measured the chemical shifts of 95 both 13C nuclei in CFSCN. These are 129.0 ppm and 130.9 ppm down field from methyl iodide for the methyl and the cyanide carbons, respectively. Using Equations (120) and (121), and also assuming that TJ can be re- lated to 10 by Hubbard's relation (Equation (28)), we obtain for each temperature two possible sets of values for C11 and CL of both the methyl and the cyanide carbons. These are listed in Table 10. Since, for cyanide carbon, the cylindrical electron cloud in the C-N triple bond should make the chemical shielding highly anisotropic, the sets with Cl] positive and CL negative seem to be more reasonable. The chemical shielding anisotropies corresponding to the chosen C values are 380 ppm and 330 ppm at -l4°C and 28°C, respectively. On the other hand, the methyl carbon should not have too large a chemical shielding anisotropy and therefore the sets of values with both C11 and Cl negative will be more resonable. Making this choice the chemical shielding anisotropies of the methyl carbon are -140 ppm and -160 ppm at -14°C and 28°C, res- pectively. In the calculations for the cyanide carbon, we have neglected the contribution from the chemical shielding anisotropy to the total relaxation rates. This can be:justified as a good assumption by using Ao~360 ppm in Equation (73a) which gives csa _ 2 2 -—w 1 l 15 ° (Ao)21'e < 0.00016 sec' . R (122) This value is about 2% of the total relaxation rate and therefore can be neglected at the temperature studied. It should, however, be noted that at lower temperatures the chemical shielding anisotropy might give a considerable contribution to the cyanide carbon relaxation rates. 96 Table 10. Spin-rotation constants of 13C in CF r (0C) C|l(kHz) cl(kHz) 13CF3 28.0 -2.575 -0.588 28.0 1.185 -1.552 .14.0 -2.470 -0.615 —14.0 1.763 -1.700 13CN 28.0 -1.721 -0.433 28.0 0.748 -l.066 -14.0 -1.593 -0.465 -14.0 1.098 -1.155 g 8There are two possible sets of values for each temperature. (See text for explanation). r-r‘ 97 6. 130, 19F Chemical Shifts and JCF Finding the trends in chemical shifts and coupling constants has been 150,195 a traditional subject for NMR studies . Muller and Carr196 have found a linear relationship between the 19F chemical shifts 6 and the F 13 _19 C F coupling constants JCF of several families of related compounds including tetrahalomethanes. DeMarco et_al,197 have also tried to rationalize the 6 6 and JC C’ F F concluded that the substitution of a less electronegative halide on one values in some trifluoromethyl halides and of the fluorines in CF4 tends to increase the 13C chemical shielding and 19 the carbon- fluorine coupling constant J but to decrease the F CF chemical shielding. The increase in JCF is evidently due to increasing multiple bonding between carbon and fluorine atoms which will also decrease the bonding polarity and therefore the chemical shielding of 19F nucleus. In Figure 17 we plot JCF versus 6F for all the fluorohalomethanes studied in this work (the parameters used are listed in Table 11). It is obvious that CFSCN lies on the same line as all the trifluoromethyl halides and is very close to CF4. This is evidently due to the smaller mass of the CN group and its greater electronegativity which make it behave somewhat like a single fluorine atom. The effect of substitution of a fluorine atom by another group X was also observed in the anisotropy of 19F chemical shielding. Montana g§_gl,192 found that the anisotropies of 19F chemical shieldings in CFSX molecules increased in the order CFSCI, -13.4:0.5 ppm < CF3Br, 2.7:0.3 ppm < CF31, 16.8:0.2 ppm (in the molecular axis systems). They also found that the component of the chemical shielding along an axis parallel to the symmetry axis of the molecule all was fairly constant 98 l l 1 1 1 l l 0 40 80 8F ("”0 Figure 17. Plot of JCF versus 6F for some fluorocarbon derivatives. Table 11. Comparison of in some fluorocarbon derivatives. 13 99 C and 19 F chemical shifts and JCF values Compound JCF(Hz) JCCF(Hz) 6C(ppm) 6F(ppm) CF3C1 299(a) 125.5(f) 33(a) CF3Br 324(a) 112.7(f) 21(a) 323 20.7 CF31 344 78.2(f) 5(a) 5.29 *CF3CN 266 129.0(d) 59.7 CFECN -56 130.9(d) CF3CC13 283(c) 120.1(f) 62(f) Cpgcc13 -43.l(c) 91.0(f) CF4 260(g) 119.9(f) 62(f) 257(a) 69(8) CFC13 337 137.7(d) 0 CFBr3 371 67.1(d) -7(e) 372(b) CF2012 325(a) 8(a) CFzBrz 358(a) -7(a) cpz=c012 289 -44.2 90 289.9(c) -43.7(c) 88.5(c) (a) Reference 196. (b) P. C. Lauterbur, "Determination of Organic Structure by Physical Methods", F. C. Nachod and W. 0. Phillips, Ed., Academic Press, New York (1962). (c) G. V. D. Tiers, J Phys. Soc. Jap. 15, 354 (1960). (d) Measured relative to CH3I. (e) c. Filipovich and G. v. 0. Tiers, J. Phys. Chem. 63, 761 (1959). (f) Reference 197. (g) E. L. Motell and G. E. Maciel, J. Magn. Resonance, Z; 330 (1972). f0) 18) sh: HE +1. pe: U 1 0): Se: me' 1‘8 re 1} 100 19 for F in these molecules. In Figure 18 we plot Ao(both in the molecu- lar axis system and in the C-F bond system) versus the absolute chemical shielding of 19 F. In view of the well-behaved trends shown in Figure 18, we can predict the chemical shielding anisotropy of 19F nuclei in CF3CN by interpolation. The Ao values obtained in this way are -48 ppm and +142 ppm for the molecular axis system and the C-F bond system, res- pectively. From the anisotropies so obtained we found for CFSCN, 0I|=215 ppm, ols263 ppm in the molecular axis system.and oll=360 ppm, olsl99 ppm in the C-F bond system. These results have been used in Section 4 for calculating the effective spin-rotation constant Ceff' Table 12 summarizes the physical constants and the motional para- meters obtained from this study of CF3CN. B. 13C and 20 Relaxation Studies and the Motional Anisotropy in Liquid CDBr3 1. Introduction As has been discussed before, for a symmetric-top molecule the reorientational motion can be completely described by observing the relaxation of two quadrupolar nuclei occupying two different sites in the molecule68. It seemed that CDBr3 would be suitable for this study but, unfortunately, the NMR relaxation rate of bromine nuclei is too fast to be measured. Therefore, we still have to obtain a second rela- tionship using the relaxation of the 13C nucleus, which is again com- plicated by contributions from several different relaxation mechanisms. Recnetly, a number of Raman spectroscopic studies of the molecular 131,174,176 motions in liquid bromoform have been reported . Bartoli and 101 100'- 40' (PPM) Figure 18. 100 0 60 89 (PPM) Plot of Act versus 5F for CF3X, X= CC13, F, Cl, Br, I, i and CN; so d circles are in the C- F bond coordinate system and open circles are in the molecular coordinate system. Tab 102 Table 12. Summary of the physical constants and the motional parameters of CF3CN. Physical Constant Value Note AEII 0.89 kcal/mole AEL. 1.40 kcal/mole CO(19F) -4.29 kHz ceff(198) -S.58 kHz c||(19p) -8.34 kHz CL(19F) -3.46 kHz 00(19F) -48 ppm a 142 ppm b Cll(13CF3) -0.60 kHz c 01(13CF3) -l.63 kHz c Cll(13CN) -l.66 kHz c Cl(13CN) 0.92 kHz c 00(13CF3) -150 ppm c 00(13CN) , -360 ppm c a. In the molecular axis coordinate system. b. In the C-F bond coordinate system. c. Average of the values at two temperatures. ti to us :10 ca me hr in US' C1. 103 Litovitz174 have analyzed the depolarized Raman lineshape of bromoform and found the correlation time of the tumbling motion relat 22°C to be 5.3:2 psec. Wright131 has also analyzed the A1 lines of the Raman 76 have spectrum and obtained a similar result. Peterson and Griffiths1 made a variable-temperature Raman spectroscopic study on this compound and obtained an activation energy AE ~1.7 kcal/mole for the tumbling motion. This activation energy can also be obtained by measuring the relaxation times of 2D in deuterated bromoform. Unfortunately, the activation energy AEl obtained from the previous NMR relaxation study131 does not agree with that obtained from the Raman work. Farrar gt_§l,203 have measured the 13C nuclear magnetic relaxation times of CHBr3 and found that the dominant relaxation mechanism was due to the scalar coupling between 13 3 C and bromine nuclei. An attempt at C and the 2D relaxation time measurements to study the 31 using the 1 motional anisotropy of CDBr3 has been made1 . We report here our carefully measured 13C and 20 relaxation rates and the motional para- meters derived from them. 2. 20 Relaxation and the Tumbling Motion The measured 20 spin—lattice relaxation rates in 97% deuterated bromoform131, which had been carefully purified and degassed, are listed in Table 13 and are plotted versus reciprocal temperature in Figure 19. Using the reported quadrupole coupling constant 177:5 kHzlgs, we cal- culate the correlation time of the tumbling motion to be 1 = 3.30x10' 3exp(1.67 kcal mole‘l/RT). (123) Tel- 104 Table 13. 2D spin-lattice relaxation rates in CDBr3. T (0C) lOOO/T (°x‘1) R1 (sec‘l) 9.0 3.55 2.93:0.13 12.0 3.51 2.73:0.12 15.0 3.47 2.62:0.14 19.0 3.42 2.67:0.14 21.0 3.40 2.58:0.09 22.0 3.39 2.57:0.11 24.0 3.37 2.43:0.10 30.0 3.30 2.37:0.08 31.5 3.28 2.34:0.10 37.0 3.22 2.17:0.12 37.5 3.21 2.07:0.09 39.0 3.20 2.22:0.09 46.5 3.13 2.02:0.10 48.0 3.12 2.02:0.12 53.0 3.07 2.03:0.08 54.5 3.05 1.98:0.10 56.0 3.04 1.92:0.07 66.5 2.94 1.66:0.08 77.0 2.86 1.70:0.09 78.0 2.85 1.52:0.08 87.0 2.78 1.53:0.08 90.0 2.75 1.52:0.07 106.5 2.64 1.40:0.05 118.0 2.56 1.29:0.08 126.0 2.51 1.24:0.06 142.0 2.41 1.15:0.10 145.0 2.39 1.14:0.06 152.0 2.35 1.17:0.04 175.0 2.23 1.01:0.05 105 2 3 4 1000/T Figure 19. Plot of the 2D spin-lattice relaxation rate in CDBr3 as a function of reciprocal temperature. thl l4 ni th ti in 1‘0 th CE 106 Our measured activation energy AE =l.67 kcal/mole agrees very well with the result from Raman work176 AElsl.7 kcal/mole. 3 3. 1 C Relaxation and Motional Anisotropy The measured 13C spin-lattice relaxation rates are listed in Table 14 and plotted in Figure 20. Usually several different relaxation mecha- nisms have to be considered for the 13C nucleus. In this compound only the scalar coupled relaxation is important. The spin-rotational relaxa- tion should be negligible since CDBr3 is much more massive than CDC13 in which only ~10% of the relaxation rate is contributed by the spin- rotational interactionlgg. The intermolecular dipolar contribution to 13 the C relaxation is always neglected since the carbon atom lies at the center of the molecule. The intramolecular dipolar contributions from 2D and the three bromine nuclei can be estimated by the following equation d 2 2 -6 2 -6 _ 8 R1 ’ ” YCcYDrCD+YBrrCBr)TB ’ 6‘10 T9 ’ (124) where we have used rcd~1.07 X and rCB 1 is about 0.003 sec' which is clearly negligible with respect to the £~1.93 R 200; with re~5 psec, R: total relaxation rates. Relaxation through chemical shielding aniso- tropy is also negligible at the low field (14.1 kG) studied; an aniso- tropy of 200 ppm results in C53 R1 7 = 2.4XIO Ta, (125) which is even smaller than Rf. 107 13 Table 14. C spin-lattice relaxation rates in CDBr3. r (°C) 1000/T (°x'1) R1 (sec-1) 167 2.27 0.72:0.07 132 2.47 0.64:0.06 101 2.67 0.55:0.05 74 2.88 0.50:0.05 49 3.10 0.45:0.05 34 3.26 0.44:0.04 22 3.39 0.44:0.05 13 3.50 0.4210.05 9 3.55 0.40:0.04 108 o 8" .0 .. .‘o 6" “.0 _ A '7 '8 o .. m " v m 4- - l 2 3 ' 4 1000/T Figure 20. Plot of the 13C spin-lattice relaxation rate as a function of reciprocal temperature. The open circles are the Ric values calculated from the 13C data of Farra§_ et a120 3; the solid circles are the R} values for C_ in CDBr3. COLL M r11 109 The existence of two bromine isotopes might complicate the scalar 13 coupled relaxation of the C nucleus. The decay of 13C magnetization Mz due to different combinations of bromine isotopes follows . 1 sc sc 3 sc M = M -— - -tR ._ - z 0{ 8 eXp( tR1,818181 1.797979)+ 3 exP( tR1,818179 SC 'tR1,817979)} ’ (126) where the second subscripts represent the three isotopic species A, B, C and 5c = sc sc sc R1,ABC R1,A + R1,B + R1,C . (127) Equation (126) implies that the 13C relaxation could be biexponential , but by a simple mathematical rearrangement of Equation (126) we can show that the 13C relaxation due to scalar coupling with the three bromine nuclei can be described approximately by a single exponential function with a time constant sc _ 3 1 - E(RSc + 85° ). (128) R 1,81 1,79 Using Equation (69) and also by noticing that Aw212<<1 for 79Br, Equation (128) can be written as 81 T ( Br) Ric = lé-{A2(C798r) T1(7gBr) + 212 81 }. (129) 4 1+Aw T1( Br) Since A(C798r)/A(C81Br) = 7(79Br)/y(818r) = 0.928 and T1(8lBr)/T1(798r) wh an w} of In 110 79 81 . . . . = Q( Br)/Q( Br) = 1.433, Equation (129) can be further Simplified to 1.71T1(798r) }. l+1.08XI014Ti(7gBr) Ric = 148.0 J2(C798r){ T1(798r) + (130) 79 79 where J(C Br)= A(C Br)/21ris the scalar coupling constant between 13C and 798r in unit of Hz and T1(798r) can be expressed by - 2 ' T11(798r) = 3.95 (Q.C.C.) T6 = 1.43Xl01819, (131) where we have used 601.1 MHz201 for the quadrupole coupling constant 7 . . . . . . . . of 9Br. re 15 the correlation time describing the motion of the princ1- pal axis of the electric field gradient tensor which can be assumed to be along the C-Br chemical bond. From Equation (111) and with 0~7l.9°200, 19 can be written as 1.57 1.84 5+r + 1+2r) - 1913(r)’ T6 = Te‘L(0.126 + (132) where r is the reorientational anisotrOpy defined by Equation (113). Since there are two unknowns in Equation (130), the exact solution is not possible without any other conditions. However, since the relvalues have been directly obtained from the deuterium relaxation studies, we can assume different values for the activation energy of the spinning 79 Rsc motion AEII and the coupling constant J(C Br) and calculate 1 values at different temperatures from Equation(l30). We then try to fit these values to those measured experimentally by varying AEII or, alternative- ly, use the measured Ric values to calculate J(C798r) by assuming different values for 08". In this way it was found that the best fit 111 of the Rsc values using Equation (130) required that AEII=O.9 kcal/mole 1 and J(c798r)=105 Hz. 203 13 . . Farrar, et al. have measured the C relaxation rates in CHBr3 at several different temperatures. By subtracting the dipolar contribu- tion from the proton in CHBr3 using our measured rel values,we find that their relaxation rates Ric are a little higher than our results for CDBr3 (see Table 15 and Figure 20). By using the same technique described above for fitting AEII and J(C798r) using Equation (130), we find that AE||~1.35 kcal/mole and J(C798r)~121 Hz in order to provide the best fit to Farrar's data. In view of the crude approximations made, and the unknown validity of our fitting method, very precise conclusions cannot be made from this study. However, we can compare our result with that for CDCl3 for which the motional parameters were very well determined by measuring the 2D 172 and 35Cl relaxation times and Table 16 summarizes these results. Since the geometrical anisotropies of these two compounds are comparable, one would expect that the anisotropies of the molecular reorientations should also be comparable and this is indeed true. Another point that can be noted is the value of the scalar coupling constant J(C798r). Shoup and Farrar199 have found J(CSSCI) in CDCl to be ~23 Hz. Since 3 J(c35c1)/J(c798r) ~ y(3501)/y(798r), this would give J(c798r) about 59 Hz which is much smaller than our eStimated values. Briguet et_al?02 have found J(C798r) in CHBr3 to be ~41 Hz, but their measured Ric at room temperature was about 0.14 sec'1 which is much lower than our value and the other reported valueslss’199 . Although a higher aniso- tropy of the reorientational motion will give a lower calculated J(C798r), we found that the lowest possible J(C798r) (about 90 Hz for AE|'=0 with * 112 Table 15. The separation of the different contributions to the total relaxation rates of 13C in CHBr3. . b c d a t d sc 1' 1000/1‘ 19L R1 R1 R1 (OK) (OK'I) (psec) (sec'l) (sec-1) (sec-1) 124 2.52 2.72 0.90 0.0653 0.835 107 2.63 2.97 0.86 0.0713 0.789 90 2.75 3.30 0.97 0.0792 0.711 69 2.92 3.82 0.70 0.0917 0.608 50 3.09 4.37 0.61 0.105 0.505 31 3.29 5.17 0.61 0.124 0.486 11 3.52 6.25 0.61 0.150 0.460 aCalculated from Equation (123). bFrom reference 203. c d- 10 R1-2.4OXI0 161: d sc _ t d R1 - Rl - R1. 113 Table 16. Physical properties of CDBr and CDCl . 3 3 Physical Property CDBr3 CDCl3 Ii (10'40g cmz) 685a 264b Ill (10’40g cmz) 13383 498b d (Q.C.C.)D (kHz) 177c 167 (Q.C.C.)x (MHz) 601e 79f O a g rCD (A) 1.07 1.100 AE k 9h 7b 11 ( cal/mole) 0.. _ 0. 1.351 AEi (kcal/mole) 1.67h 1.6b 1.73 r (at 20°C)k 3.08 1.88b 1.81 aCalculated from the geometry in Reference 200 . bReference 172. cReference 198. d 8Reference 201. f J. L. Ragle, G. Minott and M. Mokarram, J. Chem. Phys. 69, 3184 (1974). E. A. C. Lucken,"Nuclear Quadrupole Coupling Constants", Academic Press, New York (1969). gM. Jen and 0. R. Lide, Jr., J. Chem. Phys. 36, 2525 (1962). hFrom this work. iFrom the 13 jReference 176. k C data of Farrar g; 31. 203 r is the reorientational anisotropy (Equation (113)). 114 our 13C data and is ~98 Hz with Farrar's 130 data) was still high. This is probably an anomalous property of highly brominated methanes. We have not been able to discover any reasonable alteration in the data analysis method which would reduce the calculated J(C798r) by a factor of two or three§01. l9 13 C. F and C Relaxation Studies in Liquid CFBr3 1. Introduction A complete determination of the reorientational anisotropy is almost impossible in CFBr3,which does not have any observable quadrupolar nucleus. However, from the small dipole moment (0.58 D206) and the fairly spherical shape of the molecule,the molecular reorientation can be assumed to be isotropic. This has been found to be true in CFC13171 which has a dipole moment ~0.46 D and also has a nearly spherical molecular shape. With this assumption, the 19F and 13C relaxation rates in liquid CFBr3 can be analyzed to provide estimates of a few physical constants of this molecule which cannot be obtained otherwise, e.g., the scalar coupling constants between fluorine and bromine,and between carbon and bromine,and the spin-lattice relaxation times of the bromine nuclei. 19 2. F Spin-Lattice Relaxation and Isotropic Reorientation The measured 19F spin-lattice relaxation rates R; F are listed in 3 Table 17 and plotted against reciprocal temperature in Figure 21. A dominant spin-rotational relaxation is evident from the negative slope of the plot at high temperatures. At low temperatures the dominating Ta . .071623345681111 ‘ 115 Table 17.19F spin-lattice relaxation rates in CFBrS. T 1000/T R (°C) (°x‘1 (sec'l) -75.0 5.05 0.248:0.009 -67.0 4.85 0.208:0.007 -66.0 4.83 0.213:0.003 ~59.0 4.67 0.177:0.004 -51.0 4.50 0.162:0.004 -50.0 4.48 0.155:0.006 -4l.5 4.32 0.148:0.003 -42.0 4.33 0.144:0.005 -3l.0 4.13 0.133:0.002 -21.0 3.97 0.121:0.002 ~18.0 3.92 0.125:0.005 -17.0 3.91 0.122:0.002 -ll.0 3.82 0.120:0.002 -9.0 3.79 0.123:0.006 -1.0 3.68 0.120:0.003 0.0 3.66 0.125:0.002 7.0 3.57 0.130:0.006 18.0 3.44 0.130:0.002 6.0 3.58 0.124:0.001 27.0 3.33 0.138:0.004 36.0 3.24 0.155:0.009 37.0 3.23 0.142:0.002 46.0 3.13 0.156:0.004 57.0 3.03 0.170:0.006 69.0 2.92 0.188:0.006 85.0 2.79 0.203:0.006 104.0 2.65 0.233:0.007 122.0 2.53 0.259:0.009 146.0 2.39 0.319:0.016 170.0 2.26 0.370:0.014 116 0.5 — 4 I :1 )1. RHSEC O . O H 0.05 l 3 1 lOOO/TtoK) 4 5 Figure 21. Plot of the 19F spin-lattice relaxation rate as a function of reciprocal temperature; 0: R1,F3 __3 R ---: RffF. d . 1,F’ 1‘8 in DE 85 117 relaxation mechanism is probably that due to the intermolecular dipolar interaction. The other relaxation mechanisms can be easily shown to be negligible. The intermolecular dipolar relaxation rate in CFBr3 can be estimated by use of Equation (133) d "HZYgN 2 2 R = ————-'(YF + IOYBr). (133) l,F 405a where a, the molecular radius, and N, the number of molecules per unit V volume, can be estimated from the density by using Equations (108) and (109); 05’ the self-diffusion coefficient, can be estimated from the shear viscosity n of liquid CFBr3 by the equation s = 60::fr ’ (134) where fr=0.5 is a correction factor from the microviscosity theory211 Gillen gt_al,17l have reviewed several studies and found that the intermolecular dipolar relaxation rate calculated from Equation (133) did not agree with the directly measured values from the isotopic 163’164; however, the activation energies obtained from dilution method both methods were essentially the same. Therefore, instead of directly using the calculated values from Equation (133), we use the same activation energy, which was found to be 2.97 kcal/mole, and vary the pre-exponential factor so that the spin-rotational relaxation rates Rifp ( RifF=R§’F-Rg,p) at low temperatures follow an Arrhenius relation with the temperature. The best values of Rg,F and Rifp obtained by this method of fitting are listed in Table 18 together with the theoretically calculated R? F values from Equation (133). hm~0£mfi Lm~_.u.o~CELL+C., 118 .e.wm - e.w¢ u a.H .uxou oomm .Amnfiv oomoeoom eoem oooemoofioom .nmoav :owuonem seam wouofiaofiouo .memfiv cowuoaem scum pouofisofloup .mwofiv :ofiuosem seam pouofioofiouo .mousuonoAEou yoguo we mo=Hm> one seam wouomomonuxmn .flommfiu cam mmm .mxgm .ano .5 .cuxEm .m .0 one Homafiz .0 .mo ooH.o mofio.o meoo.o wwm.o eon.” ooH.n co.” e.~oH oo.n o.oo emfi.o Hefio.o Nofio.o Hoo.o oefi.~ owo.m m~.H ~.ooH o~.n o.oe eHH.o memo.o omHo.o eHo.o oee.o moo.m om.“ H.ee He.» o.o~ eeo.o NeNo.o meNo.o oNo.o emo.o meo.m oo.~ ~.oo oo.m o.o oeo.o oeeo.o a:mo.o oeo.o use.o NNo.m ooe.~ oH.eo mo.» o.o~- 15-66.: n.-...o 13-66.: 16-6.: 23-6..Nseo A“: 18.: means a.-eso hes: a H a H a He oz--oH he on e c e> e\ooo~ a 59mm wmmxoUvm Acuuv v m+ u m m .nummu new money eowuoxofiou uoHomfip noasooaoeuouefi one we mosmo> Houeoaanomxo one Hoowuouoonu on» me :omwnomaou .mH ofipoh 119 The low temperature spin-rotational relaxation rates are found to have an activation energy about 1.42 kcal/mole and can be repre- sented by air = 1.343 exp(-l.42 kcal mole'l/RT). (135) Since, at low temperatures, the molecular motion is very likely to be~ diffusive, Hubbard's relation (Equation (78)) can be used in Equation (80) so that 2 2 2 2 4w I C Rfr = §1_%I-IcgffTJ = 2 eff , (136) n 3” Te where the symbols have their usual meanings. The 19F chemical shift in CFBrs was measured and found to be -7 ppm relative to CFC13. From Equation (116) we calculate the spin-rotation constant Co to be ~l.40 kHz. Using this value for Ceff in Equation (136), we find the reorientational correlation time re to be expressed by the equation 13exp(1.42 kcal mole'l/Rr), (137) T9 1.62X10' which gives 1621.75 psec at room temperature. Miller and Smyth204 have measured the dielectric relaxation times TB of liquid CFBrS at four temperatures. These are shown in Figure 22 together with the re values from Equation (137). The ratio TE/Te in the temperature range where TB is available is found to be about 3.4, which is very close to the theoretically predicted value 3.0 for 120 -12 N 10 SEC 1000/ T(°K) Figure 22. Plot of the reorientational correlation times as functions of reciprocal temperature for CFBr3; —: Te; --‘: TE. 121 molecules undergoing diffusive reorientation. 3. 19F Spin-Spin Relaxation and J(F798r) The measured 19F spin-spin relaxation rates R2,F are shown in Table 19 and in Figure 23. In addition to the intermolecular dipolar interac- tion and the spin-rotational interaction, the scalar coupling between 19F and the bromine nuclei also gives a considerable contribution to the spin-spin relaxation rates. Therefore the scalar-coupled relaxation rates can be obtained by measuring the differences between R5 and R} as shown in Figure 23. From Equation (70), and since szTi>>l, the spin-spin relaxation rates due to the scalar coupling between 19F and the three bromine nuclei in CFBr3 can be written as 83° = 7.5 «2( J2(F798r)T1(798r) + J2(F818r)T1(818r). (138) This equation can be further simplified by using the relations: J(F798r)/J(F818r) = y(798r)/y(818r) = 0.927 and T1(818r)/T1(798r) = ‘(79 Q Br)/Q(8lBr) 2 1.433. Therefore, Ric = 200.6 J2(F798r)T1(7gBr), (139) where T1(798r) is given by 79 TI1( Br) = 3.95 (Q.C.C.)2 t9. (140) Table 19. 122 19F spin-spin relaxation rates in CFBr3 7 (°C) 1000/T (°K'1) R2 (sec‘l) 32.0 3.28 0.195:0.008 8.0 3.56 0.179:0.017 18.0 3.44 0.183:0.016 49.0 3.10 0.224:0.012 -78.0 5.13 0.340:0.023 -68.0 4.88 0.257:o.024 -54.0 4.57 0.204:0.010 -56.0 4.60 0.204:0.009 .37.0 4.24 0.176:0.008 .24.0 4.02 0.16l:0.009 -l6.0 3.89 0.157:0.011 -3.0 3.70 o.173:0.011 -5.o 3.73 0.158:0.008 55.0 3.05 o.24o:o.005 64.0 2.97 0.278:0.012 85.5 2.79 0.288:0.015 106.5 2.64 0.40310.015 144.5 2.40 0.500:0.048 (108love? o. 123 0.5 14:13. I 41 z 0.1 "' G 0.1 -‘ TA 0 L) “U 33 ~ 8 o: C 0.05- . - O O 'F C t ‘P 0.02 “ J“ " 1. 1 1 3 IOOO/T(°K) 4 Figure 23. Plot of the 19F spin-spin relaxation rate in CFBr3 as a function of reciprocal temperature; 0 :Rt ; . Rsc 2,F 2,F° 124 The quadrupole coupling constant for 81 found to be 518.96 Mizzo1 in the solid phase. The corresponding quadru- Br in CFBr3 has been measured and pole coupling constant of 798r is therefore ~621 MHz. From Equations (139) and (140) we obtain Ric = 1.32X10 16.179202 Br)/Te. (141) Using re in Equation (137) and the measured Rgc, we find the coupling constant J(F798r) to be ~30.5:l.4 Hz. This can be compared with the J(F3 SCl) value (~11. 9+0. 4 Hz) in CFC13 obtained by Gillen et a1171. The ratio of J(F798r)/J(F35 Cl) is found to be ~2.56 which is in excellent agreement with the ratio of the gyromagnetic ratios of 798r and 3501 (7(798r)/y(3SC1)=2.55). 4. 13C Spin-Lattice Relaxation and J(C798r) 13 It has been found that the scalar coupling between C and bromine nuclei is a dominant relaxation mechanism in highly brominated methanes. This is also true in CFBr3. By the same approach used in analyzing the 13C relaxation rates in CHBrs, we find that the only two important mechanisms contributing to the 13C relaxation in CFBr3 are the dipolar interaction and the scalar coupling due to the three fast-relaxing bromine nuclei. The dipolar relaxation rate R? C can be estimated by the following equation: 4 2 R1,C W” (Yrrcr * ISYBr rCBr)T9’ (142) 125 where we have assumed that the reorientation of the molecules is iso- tropic so that the same correlation time T6 can be used to describe the reorientation of both the C-F and the C-Br vectors. Table 20 summarizes the experimental data (values of RE,C at various temperatures) along with the calculated values of R?,C; the scalar-coupled relaxation rates Rifc obtained from R§,C"R?,C are also given and are plotted in Figure 24. By using Equations (129) and (140), we can then calculate the coupling constant J(C798r). This was found to be ~124.2t7.0 Hz which is comparable to that in CDBr3. The various quantities used in the calcu- lation of J(C798r) and T1(798r) are summarized in Table 21. 5. Extended Diffusion Models The spin-rotational relaxation rates can be used to obtain the angular momentum correlation times TJ. From Figure 21 we found that sr Rl,F temperatures. This is evidently due to the increasing importance of plotted versus 1/T deviates from the Arrhenius equation at high the inertia effect. In Figure 25, we have plotted the reduced re- orientational correlation times 1; versus the reduced angular momentum correlation times 13. It is not surprising that the points at low temperatures lie so close to the curve predicted by Debye's diffusion model since we have used this model to derive the we values. At higher temperatures, the points are approaching the J-diffusion curve. We conclude that this is another proof of the validity of the extended J-diffusion model in describing the reorientations of the molecules in the liquid state. Table 22 summarizes the physical constants used or obtained in this study. 126 Table 20. Summary of the 13C spin-lattice relaxation rates in CFBr 3. T lOOO/T a at Rd b RSCC Te 1.C 1.C 1,c (°C) (OK-1) (psec) (sec-1) (sec-1) (sec-1) -73 5.00 5.77 0.56:0.10 0.0249 0.535 ~56 4.61 4.37 0.58:0.17 0.0189 0.561 -53 4.54 4.16 0.59:0.11 0.0180 0.572 -35 4.20 3.26 0.62:0.10 0.0141 0.606 -8 3.77 2.40 0.69:0.11 0.0104 0.680 27 3.33 1.75 0.84:0.13 0.0076 0.832 55 3.05 1.43 1.00:0.15 0.0062 0.994 94 2.72 1.13 1.45:0.15 0.0049 1.345 3Calculated from Equation (137). bCalculated from Equation (142). c sc _ t d R1,C- R1,C'R1,Co 0.5 .0 N 0 SEC 0.05 0.02 QOI 0.005 127 “- “if-mi: .J .4 l I l 3 4 5 IOOO T(°K) Figure 24. Plot of the 13C spin-lattice relaxation rate in CFBr3 : R¥,C; ---: Rifc. as a function of reciprocal temperature; 0: Rt 1,C‘ 128 Table 21. Summary of the information used in the calculation of 79 T1( Br) and J(C7gBr) in CFBr3. T IOOO/T x: T1(79Br)b Rice J(C798r)c (°C) (°x'1) (psec) ( sec) (sec-1) (Hz) -73 5.00 5.77 0.114 0.535 136.0 -56 4.61 4.37 0.150 0.561 129.7 —53 4.54 4.16 0.158 0.572 129.2 -35 4.20 3.26 0.201 0.606 124.2 -8 3.77 2.40 0.274 0.680 118.8 27 3.33 1.75 0.375 0.832 116.4 55 3.05 1.43 0.459 0.994 116.8 94 2.72 1.13 0.581 1.345 122.3 Ave: 124.217.0 8Calculated from Equation (137). bCalculated from Equation (140). cCalculated from Equation (130). 129 .93 "my". "mummu new 2:: coflumaouuou asucosos amasmcm convoy mamuo> oswu :ofiumflonnoo Hm:0fiumpcoauoon vousvon mo uon .mm ohzmwm N . _ md .9 Nd .6. mod — .. .../ a _ I 4 _ a — 130 Table 22. Physical properties of CFBr3 used or obtained in this study. Physical Property Value Reference 81 (Q.C.C.)( Br) 518.96 MHz 201 -8 rCF 1.44XIO cm 205 rCBr 1 .91><10'8 cm 205 II I 1334x10'40g cmz Ii. 785X10-4Og cm2 H 0.58 D 206 cabs 188.6 ppm Co -1.40 kHz 79 J(F Br) 1(30.Stl.5) Hz J(C79Br) i(124.2i7.0) Hz 131 13 D. C Nuclear Magnetic Relaxation and J(C79Br) in Liquid CHQBr7 1. Introduction In order to obtain further information concerning the scalar coupling constant between carbon and bromine nuclei which cannot be directly observed in NMR spectroscopy, we have chosen a less highly brominated methane, methylene bromide, whose motional parameters have been obtained recently by Sandhu207 so that the separation of different mechanisms contributing to the 13 207 C spin-lattice relaxation rates can be achieved. Sandhu studied the relaxation of 1H and 2D in CHZBrz and obtained the rotational diffusion constants parallel and perpendicular to the C2 symmetry axis of the molecule. From Figure 4 of Sandhu‘s paper, and using his best estimated deuterium quadrupole coupling constant, 181 kHz, we obtain 1 DJ. = 2.19x10 2exp(-2.23 kcal mole‘I/RT) (143) and 0" = 9.07x1012exp(-2.09 kcal mole'l/RT). (144) Gordon gt_gl,208 have also studied the proton relaxation of CHZBr2 in CS2 solution and found that the spin-rotation interaction gave a con- siderable contribution to the proton spin-lattice relaxation rate in dilute solutions of CHZBr2 at high temperatures. They also found the reorientational correlation time of the H-H vector to be ~0.74 psec at room temperature which is only about one-half the value 1.54 psec 132 found in neat CHZBrz. Also, Sandhu's relaxation studies in neat CHZBr2 showed no contribution of the spin-rotational relaxation. Evidently the intermolecular interaction among the CHZBr2 molecules has an im- portant effect on the reorientational motion of the molecules. Wang and 2 his co-workers 09 have also investigated the molecular motions of CHZBr2 by depolarized Rayleigh and Raman scattering experiments. Several con- clusions were reached from their studies. First of all, they found that the reorientational correlation time obtained from the Raman lineshape analysis was not equal to the single-particle reorientational correlation time, which was found to be 3.4 psec from the depolarized Rayleigh scattering experiment at room temperature. Second, from a variable temp- erature study of the depolarized Rayleigh scattering in neat CHZBrz, they obtained an activation energy ~2.8 kcal/mole for the tumbling motion of the molecules. Third, their plot of the correlation times versus n/T showed a slope about ten times smaller than that predicted by the Stokes-Einstein relation but was in agreement with that predicted by 210 considering Hu and Zwanzig's slip-boundary condition and modeling CHZBr2 as an ellipsoid. Wang's result20 9 for the room temperature reorientational corre- lation time, 3.4 psec, was in very good agreement with Sandhu's result (teie3.5 psec at room temperature) but the activation energy of the tumbling motion was about 25% higher than that obtained by Sandhu ( ~2.23 kcal/mole). Nevertheless, since Sandhu's studies covered a wider 13C relaxation rates will be temperature range, the analysis of our based on the reorientational diffusion constants expressed by Equations (143) and (144). 133 13 2. Mechanisms of C Relaxation and J(C79Br) Among various mechanisms of 13C spin-lattice relaxation, only the dipolar interaction with the two protons and the scalar coupling with the two fast-relaxing bromine nuclei should be considered. The other mechanisms can be easily shown to be negligible. The dipolar relaxation rate can be expressed by R?,C = 2yéy§fi2regtfi = 5.37x1olorH, (145) where TH is the reorientational correlation time of the C-H inter- nuclear vector which can be estimated from the geometry of the molecule and by using Equations (111), (143) and (144). In Table 23, we summarize the calculations of R$,C and therefore the scalar-coupled relaxation rates Rifc,which is the only other important relaxation contribution and can be obtained by subtracting Rg’c from the measured total relaxation rates R1,C' The results obtained by these calculations are also plotted as a function of temperature in Figure 26. The extracted scalar-coupled relaxation rates can be used to estimate the coupling constant J(C79 Br) by the same technique used before (see Section B). For this molecule, which contains only two bromine nuclei, we obtain sc R1 0 = 10n2J2(C7gBr)T1(7gBr)(1+ 1.71 1+1.02x1014Ti(7gBr) ) (146) and T1(79Br) = 3.95(Q.C.C.)21 (147) Br’ 134 Table 23. Sungry of the quantities employed in the calculation of the C dipolar and the scalar-coupled relaxation rates in CH Br . 2 2 t a db scc T 1000/T R1,C TH R1,C R1,C o - - - - ( C) (OK 1) (sec 1) (psec) (sec 1) (sec 1) -47.5 4.42 0.33710.03S 4.93 0.265 0.072 -29.5 4.11 0.275:0.030 3.54 0.190 0.085 -5.5 3.74 0.240:0.025 2.38 0.128 0.112 28.0 3.33 0.24Si0.020 1.54 0.083 0.162 58.0 3.02 0.276t0.021 1.10 0.059 0.217 83.0 2.81 0.340i0.040 0.88 0.047 0.293 at" = 4.33x10‘14exp(2.13 kcal mole'I/RT). bCalculated from Equation (14S). sc Rt _ Rd c 8 R1,c 1,0 1,0' 135 0.5 - mp t E 6 R1 é) 0.2 *- _ T o sc ~‘ 8 R1 cf 4‘ \ _. 001 fl — d R] .. 0.05— ._. dL I 1 1 l 3.0 3.5 4.0 4.5 1000/ T Figure 26. Plot of the 13C spin-lattice relaxation rate in CHZBrZ as a function of reciprocal temperature. 136 where TBr is the reorientational correlation time of the principal axis of the electric field gradient around the bromine nuclei which is assumed to be along the C—Br bond. The other symbols have their usual meanings. Again, far can be obtained from the geometry of the molecule and by using Equations (111), (143) and (144). From Equations (146) and (147) and the extracted scalar coupled relaxation rates, we calculate the coupling constant J(C79Br) to be m58.7il.4 Hz. Table 24 gives a summary of the quantities used in this calculation. 3. Discussion Of the three brominated methanes studied, we found that only methylene bromide has a scalar coupling constant J(C7gBr) (58.7 Hz) comparable to the coupling constant J(C3SC1) (23 Hz) obtained by Shoup and Farrarl 99. The values of J(C798r) in both CDBr3 (121 Hz or 105 Hz) and CFBr3 (124 Hz) seem to be too high if the relation JCX/JCY=Yx/YY holds approximately. Although Freeman gt_§l?12 have estimated the magnitude of J(CSS C1) to be ~49 Hz from double resonance (proton) line profile studies in connection with 35Cl wideline measurements of CHC13, it was argued by Shoup and Farrar that Freeman's result was possibly in error for many reasons. If Freeman's result for J(C35 Cl) was used to estimate J(C79Br), we find the magnitude of J(C79Br) is about 125 Hz which is very close to our results obtained for CDBr3 and CFBrz. However, since the coupling constant J(C79Br) in CHzBr2 was obtained by the same technique, we suspect that other factors should be taken into account to explain this large discrepancy. It has been well known that increasing the 0 bond electron density between carbon and halogen atoms tends to increase the coupling 137 Table 24. Summary7 of quantities used in the calculation of T1(F7 9Br) and J(C 7gBr) for CHzBrZ. T 1000/T 10100 a 10100 b rgr T1(79Br) Rife J(C79Br) (°C) (OK-1) (cmzsec'1)(cm2sec'1)(psec) (psec) (sec'l) (Hz) -47.5 .43 8.59 .52 5.10 0.138 0 072 60.5 -29.5 .11 12.0 .17 3.62 0.194 0.085 57.5 -5.5 .74 17.7 .29 2.42 0.290 0.112 57.8 28.0 .33 27.3 .22 1.55 0.453 0.162 58.1 58.0 .02 37.8 .39 1.11 0.634 0.217 57.8 83.0 .81 47.2 .35 0.88 0.796 0.293 60.3 Ave: 58.711.4 3Calculated from Equation (144). bCalculated from Equation (143). chr = 4.30 exp(2.l4 kcal mole'l/RT). dCalculated from Equation (147). eCalculated from Equation (146). 138 150 constant JCX . Therefore, it is very likely that in tribromosubsti- tuted methanes there is an important carbon-bromine double bond 295 has also found that the substitution of a more character. Pauling electronegative atom will increase the tendency to form C=X double bonds in halomethanes. Since bromoform has one bromine atom substituting for a less electronegative hydrogen atom of methylene bromide, a higher 9Br), C=Br double bond character,and therefore a larger value of J(C7 should be expected. By the same reason, the substitution of a fluorine atom for the hydrogen atom of bromoform should further increase the magnitude of J(C79Br), which is exactly what we observed. E. Anisotropic Reorientation and 19F Relaxation in CF7CC19 1. Introduction Due to the complications of describing the molecular motions in asymmetric-top molecules, very few such compounds have been studied.‘ Wallach and Huntress213 have attempted to resolve the anisotropic reorientational diffusion constants in N,N-dimethylformamide by observing the relaxation times of three quadrupolar nuclei at different sites of the molecule. They found that rotation about an axis on the molecular plane and perpendicular to the dipole moment vector was at least thirty times slower than rotation about the dipole moment vector and that rotation about the other principal axis proceeded at rates that varied from five times (at room temperature) to twenty times (at 390°K) faster than the slowest rotation. This can be easily explained by considering the nearly planar geometry of the molecule and its large dipole moment (3.91 D). Rotation about the axis parallel to the dipole 139 moment does not destroy the alignment of the molecules due to the dipolar "head-to-tail" attraction and only expels the molecules on top and at the bottom of the molecular plane. 0n the contrary, to rotate the molecule about an axis in the molecular plane and perpendicular to the dipole moment not only destroys the dipolar alignment but also needs a large free volume, therefore, it experiences the greatest interference. The rotation about the other axis, which is perpendicular to the molecular plane, does not need a large free volume but also needs to counteract the dipolar interaction,and the interference should not be so great as that experienced by the above rotation. In CFZCCIZ, we expect that the reorientational motions should also be very anisotropic. Unfortunately, there are not enough quadrupolar nuclei in the molecule to give the number of conditions needed to completely elucidate the molecular motions. Therefore, we only attempt- ed to investigate the molecular reorientation about the F-F inter- nuclear vector from our 19F relaxation studies. 2. Molecular Motions in the Solid Phase The measured 19F spin-lattice and spin-spin relaxation rates are listed in Tables 25 and 26 and are plotted versus l/T in Figure 27. From the plots of the spin-lattice relaxation rates in Figure 27, we found that in the solid phase the 19 F relaxation was largely by the intramolecular dipolar interaction and the chemical shielding aniso- trOpy mechanisms. The intramolecular dipolar relaxation rates can be expressed by 140 Table 25. 19F spin-lattice relaxation rates in CF2=CC12. T (°C) 1000/T (°K‘1) R1 (sec-1) 30.0 3.30 0.103:0.001 18.0 3.44 0.0932:0.0029 6.0 3.58 0.0872:0.0011 -ll.S 3.82 0.079210.0016 -l7.0 3.91 0.0764:0.0024 -25.0 4.02 0.075110.0038 —19.5 3.94 0.082910.0049 -27.S 4.07 0.073210.0024 -35.0 4.20 0.0738:0.0021 -44.0 4.37 0.071210.0024 -S4.0 4.57 0.075510.0015 -66.0 4.83 0.08410.001 -75.5 5.05 0.084:0.003 -82.5 5.26 0.0908:0.0024 -88.0 5.40 0.102:0.002 -9S.0 5.62 0.114:0.001 ~100.0 5.78 0.126:0.001 -106.5 6.00 0.15310.002 -ll6.0 6.37 0.229:0.005 -119.0 6.49 0.23610.002 -123.0 6.67 0.27010.003 -124.0 6.71 0.283:0.009 -130.0 6.99 0.356:0.008 -l35.0 7.25 0.834:0.019 -l41.0 7.63 1.49610.056 ~145.0 7.81 2.19:0.22 -149.0 8.06 4.57iO.63 ~121.0 6.58 0.283:0.003 19 Table 26. F spin-spin relaxation rates in CF 141 2=Cc12. T (°C) IOOO/T (°x'1) R2.(sec-1) -130 6.99 1.093:0.097 -118 6.41 0.225:0.o12 -100 5.78 o.192:0.009 -95 5.62 o.192:0.010 -75 5.05 0.18910.017 -63 4.76 0.242:0.024 -42 4.33 o.312:0.031 -23 4.00 0.386:0.035 9 3.55 0.514:o.041 27 3.33 0.567:0.073 142 5 r— ........ .» R1 - —'- (3 R2 f ---- 0 R2- 12, 2 — .1" .7” 0!: 1 - of as 9:9. {1‘ \ rn;p. ,' or ‘\\€'\ 0 : f. U 'x l .' o G’ ' f m \\ o°\.‘ : ’ GK \\ ~51 ’W*V.‘r Q2_ \ \ ~. .’.’.”0. o\ O ~._.G.o . : \ x \\ .x" ,4. ‘ (l\ 3" 0.141., x .- 'N.. “A.\ s. .. .9 . ‘ o ”Q‘w \ ‘ \\\ \ 0.05— X 0\\ \ \ l l l \ l l 4 5 6 7 8 IOOO/T Figure 27. Plot of the 19F relaxation rate for CF2=CC12 as a function ' of reciprocal temperature. 143 d __ 3 4 2 -6 R1 "' '2- an rFF £6290): (148) where f(n,D) has been defined in Equation (88) for a planar asymmetric- top rotor. In this case, since the internuclear vector of two fluorine nuclei is perpendicular to the symmetry axis of the molecule, Equation (88) can be simplified to f(n 0) --3—(0 +0) (149) ’ ‘40R y s’ where DR=3(DXDY+DyDz+Dsz) and D5: %{Dx+Dy+Dz)' The 2 axis is defined to be perpendicular to the molecular plane, the x axis is defined to be along the C2 symmetry axis and the y axis is perpendicular to the x and z axes. From the recently determined molecular geometry214, we obtained rFF22.18 R and therefore (1 D +0 R = 4.68XI09—%-—s- (150) 1 R The relaxation through the chemical shielding anisotropy can be estimated by Equation (73a). For Ao=200 ppm this contribution will be about 10% of R3. Since the actual A0 in this molecule is not known, we will assume that this contribution is negligible. Since we only have one equation and there are three unknowns,it is impossible to solve the problem without any approximations. We can assume that in the solid phase of CFZCCl , as was found213 for N,N-dimethyl- z-Dx>>Dy’ D formamide. Then Equation (150) can be further simplified to give Z 144 Rd 2 4.68x109——-—l————-= 2.60X108/D , (151) 1 9(0y+02) .L where DL is defined as the average diffusion constant for the tumbling motion. By using Equation (151) and the measured 19F spin-lattice relaxation rates in the solid phase at low temperatures, we can cal- culate this average diffusion constant D1: The expression best fitting the experimental data was found to be 0le (9.0:3.0)x1016exp(-5.22 kcal mole'l/RT), (152) where the large uncertainty arises from having made the approximation Dx>>D Equation (152) gives DisS.ZX109 cm2 sec‘l, or Tele32.0 psec, y’Dz' at a temperature just below the melting point. But when the dipolar relaxation rate is extrapolated to room temperature Telf0.01 psec, which is unreasonally small. This implies that the reorientational motion in the liquid and solid phases should have different mechanisms. The activation energy of the reorientational motion in the liquid should be smaller than 5.22 kcal/mole since the onset of translational motion has considerably eased reorientation of the molecules. 0n the other hand, the frequent collisions in the liquid phase should slow down the reorientational motion of the molecules. In the solid phase the molecules are fixed in lattice sites and therefore reorientation is more likely to be via random jumps among different orientations of the molecules. 145 3. Rotational Motion in the Liquid Phase The analysis of the 19 F relaxation in liquid CFZCCl2 is even more difficult because several possible relaxation mechanisms are to be considered and sufficient physical constants are not available.. However, by measuring the differences between the 19F spin-spin and spin-lattice relaxation rates we obtained the scalar-coupled spin-spin relaxation rates,which followed the Arrhenius equation with temperature very well. The activation energy was found to be ~1.83 kcal/mole. The existence of two chlorine isotopes might complete the problem. The decay of the magnetization due to the scalar coupling relaxation mechanism can be shown to be I! t = 9 sc _9_ 5° .1. 5° gg' ig'exP('tR2,3535) * 16 exP(‘tR2,3537) * 16 exP(‘tR2.3737)’ (153) where Rsc - Rsc + Rsc (154) 2,AB ‘ 2,4 2,B and each ch is given by a standard expression of the form of Equation (70). Equation (153) implies a nonexponential decay of the transverse magnetization. However, a simple calculation shows that this non- exponentiality can be ignored if the relaxation rates are determined from the initial decay of the transverse magnetization. Therefore, by noticing that szTiS>>l, the scalar-coupled spin-spin relaxation rates can be expressed by 146 83° = 98.7 12(935c1)T1(3501), (155) where J2(F35C1) is the mean square value of the cis and trans coupling constants. Tl(35Cl) can be written as Til(35Cl) = 3.95(Q.C.C.)ZT¢, (156) where T¢ is the correlation time describing the reorientation of the C-Cl bond, which is assumed to be the principal axis of the electric field gradient tensor at the chlorine nucleus. T¢ can also be written as a function of the diffusion constants as in Equation (88). Since the scalar coupling constant J(FSSCl) has rarely been measured, we will try to estimate it from the magnitude of JFF in some 217 215 analogous compounds . Tiers and Kaiser216 have both measured the coupling constants JFF in cis and trans-1,2—dif1uorodichloroethylene. Their measured JFF(cis) and JFF(trans) values were 37.9 Hz and 130 Hz, respectively. By noticing that the coupling constants JAx and JAY are usually related to the gyromagnetic ratios YX’YY by JAX/JAY=YX/YY’ we can estimate J(F35C1)(cis) and J(F35Cl)(trans) to be m4 Hz and ~14 Hz, respectively. The quadrupole coupling constant of 35C1 in CFZCCIZ is not 21 available, but since Q.C.C.(35Cl) is ~73.7 MHz in cis-CHC1=CHF 8, ~70.5 m: in cis-CHC1=CHC1219 and v»76 MHz in CF2=CFC1294, we can estimate that it is ~75 Mix in CF2=CC12. Therefore, Equations (155) and (156) can be combined to give 13 R = 9.4ZXIO- /T¢. (157) 147 From the measured kgc we find 1 = 7.8SXI0' 4exp(l.83 kcal mole-I/RT), (158) ‘4 which gives T¢=1.72 psec at room temperature. F. 19F Relaxation Mechanisms for CF31 Trifluoromethyl iodide is a symmetric-top molecule with dipole moment 021.0 D220 directed along the C3 symmetry axis and a fairly an- isotropic molecular geometry (Il/I||=3.7)221. Therefore, its re- orientational motions are expected to be very direction dependent. Unfortunately, the relaxation time of iodine is too short to be measured directly. Hence, information about the tumbling motion has to be obtained by other spectroscopic techniques, e.g., a variable temperature study of the depolarized Raman or Rayleigh scattering. We have measured the 19 F spin-lattice relaxation rates in liquid CF3I,which are listed in Table 27. The separation of various relaxation mechanisms has been found to be difficult. From our measured densities and the self-diffusion coefficients (Table 28),we estimate the inter- molecular dipolar relaxation rates according to Equations (107), (108) and (109), and subtract them from the total spin-lattice relaxation rates. The results are shown in Table 29 and in Figure 28. It can be seen from the plot in Figure 28 that spin-rotational relaxation is the dominant relaxation mechanism over the whole liquid range. This has made estimation of the intramolecular dipolar relaxation rates, and therefore further separation of various contributions, impossible. 148 19 Table 27. F spin-lattice relaxation rates in CF31. o O '1 -1 T ( C) 1000/T ( K ) R1 (sec ) 3.0 3.62 0.558:0.007 4.0 3.61 0.559:0.012 -5.0 3.73 0.497:0.008 -15.0 3.88 0.48210.007 —25.0 4.03 0.435:0.007 -41.0 4.31 0.38510.004 -52.0 4.52 0.400:0.006 -63.0 4.76 0.359:0.004 -74.0 5.02 0.35410.004 -84.0 5.29 0.383:0.003 -94.0 5.59 0.37210.004 -103.0 5.88 0.37810.003 -111.0 6.17 0.420:0.006 -125.0 6.76 0.52910.005 -130.0 6.99 0.553:0.008 -137.0 7.35 0.62310.006 -135.0 7.24 0.146:0.012 -l36.0 7.29 0.134:0.012 -131.5 7.06 0.159:0.011 -127.0 6.85 0.284:0.007 -125.0 6.76 0.489:0.006 -124.0 6.71 0.45410.005 -122.0 6.62 0.459:0.003 -117.0 6.41 0.40910.004 ~109.0 6.10 0.380:0.004 -102.0 5.85 0.381:0.002 -93.0 5.56 0.369:0.003 -85.0 5.32 0.355t0.003 149 Table 28. Self-diffusion coefficients of CF31. T 1000/T g 10 05 (°C) (OK-1) (G cm-l) (cmzsec‘l) 3.0 3.62 161 2.02 -S.0 3.73 168 1.82 -15.0 3.88 178 1.57 -25.0 4.03 191 1.23 -41.0 4.31 215 0.97 -52.0 4.52 235 0.76 -63.0 4.76 260 0.60 -74.0 5.02 290 0.43 -84.0 5.29 325 0.332 -94.0 5.59 - 369 0.262 -103.0 5.88 419 0.223 -111.0 6.17 471 0.186 -118.0 6.45 531 0.122 ~125.0 6.76 600 0.183 -130.0 6.99 670 0.070 . H u H n H HoucH.um um nocuomo .Hach soHumsum aoam wouaHsonoe .HmoHv noHumocm aoum voumHsuHmuo .nmoHV :oHumsdm Seam poumHsonun .xuo: anu scum essHou «Hg» :H noaHe>m 150 moH.o owe.o eHm.o oH.o o¢.~ ma.~ ac.~ ma.o o.m~H- omH.o ooe.o «Hm.o 6H.o ev.~ ma.~ mo.~ ~n.o o.mHH- mNN.o owm.o mmH.o o~.o Nv.~ om.~ ~6.N ma.m o.moH- mo~.o mmm.o mmmo.o Nm.o mm.~ mm.~ mm.~ cv.m o.om- vom.o mmm.o NHmo.o om.o ~m.~ mm.~ Nm.~ Nm.v o.oe- wvm.o omm.o mHmo.o mm.o m~.~ am.~ ve.~ we.e o.om- «on.o uo~¢.o oc~o.o «HN.H 6H.~ om.~ amn.~ HH.¢ o.cm- HH-uomv HH-666V HH-oomV HH-umm~aov Hn-EUV HMV Hm-ao my HH-goV Hues H H H m ouugpom pm cumucH.um amoH oz--oH pa 6 H\oocH a .Hmmu :H moumu :oHpmmeoH ooHuumHuchm m map a» wcoHunnHHucou one we :oHuanmmom .mm oHpab mH 151 #- l 1 L 3 4 5 6 7 IOOO/T Figure 28. Plot of the 19F spin-lattice relaxation rate in CF31 as a function of reciprocal temperature. 152 However, the small activation energy of the spin-rotational relaxation rate (~0.6 kcal/mole) implies that this relaxation mechanism is pro- bably due to the spinning motion of the molecules as has been fbund for 13 methyl iodide by a study of the C relaxation164. II. NMR Studies of Molecular Motions in Plastic Crystalline Fluorochlorocarbons A. Introduction The possibility of molecular rotation in crystal lattices seems to have been first raised by Simon222 to account for certain solid— solid transitions and this idea was developed by Paulingzzs. Several years later Timmermans224 found a great number of molecules which exhibited analogous physical properties such as being soft and waxy, having high melting points and small liquid ranges, having high vapor pressures and subliming easily, having low entropies of fusion and high entropies of transition, etc. Such compounds were defined by him as "globular molecules" and are known as "plastic crystals" nowadays. The molecules generally have a high symmetry, or can achieve a Spherical envelope by rotation, and are believed to be rotating in the plastic crystal phase. Many experimental methods have been used to investigate the detailed structure and the dynamic properties of these molecules and crystalszzs. Most of the earlier investigations were thermodynamic studies in which the enthalpy or entropy change due to molecular motions,and the temperature dependence of the heat capacity,were 153 measured226 . The transition of a rigid solid into a plastic phase can be observed by the sharp rise in the heat capacity which occurs at the onset of molecular rotation due to the extra amount of heat needed to overcome the rotational energy barrier. A similar phenomenon has also been found in studying the dielectric constants”? of these compounds. A rapid increase of the dielectric constant signals the onset of a solid-solid transition at which the molecules begin to reorient. In a 228, it was found that higher- crystallographic study of plastic crystals order reflections could hardly be seen because of their low intensity and the fuzzy background which was caused by the randomly displaced atoms or groups with respect to the mean geometric planes of the crystal lattice. Many plastic crystals have been found to have a cubic crystal structure and, in addition, the distance between molecular centers in the plastic crystals has been found to be shorter than the maximum diameter of the molecule. Some cooperative fluctuation in the positiom of neighboring molecules therefore appears to occur when a molecule undergoes reorientational motion. The possibility of self-diffusion in plastic crystals was shown by the high mechanical mobility229 in the high-temperature plastic phases compared to that in the low-temperature phases of some crystals, and later also by radiotracer technique5230. Among all possible experimental methods, nuclear magnetic resonance spectrosc0py was found to be the most powerful one for inves- tigating the nature of plasticity in molecular crysta15231. Details of the molecular motions can be studied by both NMR lineshape analysis and 232-261 by the relaxation behavior of the nuclei . We will discuss some basic theories which are to be applied to our plastic crystal studies 154 in the following sections. B. NMR Theories Relating to Molecular Motions in Plastic Crystals 1. Linewidths Consider two spin k nuclei in a rigid lattice, the Hamiltonian describing the dipole-dipole interaction can be written as + 1 3,52 361-3102..) JC = -{ - ), (159) d 2 3 5 I‘ r + + . . where "l and 02 are the magnetic dipole moments and I is the inter- nuclear distance. From perturbation theory the eigenvalues can be found to be (for "1:“2) En = -2nuHo + uzr’3(3c0520-1)(2-3n2). n=0, 1. (160) Transitions are allowed between two adjacent energy levels with energies AB = 20{ Ho 1 g-ur'3(3cosze-1)}, . (161) which should give two separate peaks in the spectrum. When there are n spins all interacting with each other, these two peaks will be broadened out to become a Gaussian envelope. As one might expect, any kind of molecular motion in the lattice tends to modulate this dipolar inter- action and therefore the resultant lineshape of the spectrum. Bloembergen, Purcell and Pound52 considered the effect of stochastic thermal motions upon the NMR absorption linewidth and obtained a simple 155 relation between the linewidth Av and the correlation time of the motion Tc as follows: 2 . _ (Av)2 = C2 E-tan 1(2TCAV), (162) where C is the linewidth of the rigid lattice spectrum. Kubo and Tomita262 have introduced a quantum statistical formula- tion into the NMR theories and have derived a similar equation for the linewidth, nAvr 4-ln2 2 -1 c n C tan (4-ln2)° (163) (Av)2 = Gutowsky and Pake263 have also modified the Bloembergen-Purcell- Pound equation and have obtained the following equation for the correlation time TC! _ 8'ln2 ((4v12-821} C - ZwAv 2(C2-B2) T tan { (164) where B is the linewidth after the motional transition. Recently, Hendrickson and Bray264 have used a simple model of molecular motion which is based on the partition of two states, non- excited and excited, and have considered the linewidth as an arithmetic average of those in these two states. They obtained E ) = - —§-— ln( 1-- -l-). (165) ln( —l-- 1- Av C kT B Av As we shall see, these equations give estimates of the correlation 156 time of the same order of magnitude, and also give the same activation energy. 2. Second Moments of NMR Absorption Lines Theoretical calculation of the second moments of NMR absorption lines has been successfully done by Van Vleck26s. For a polycrystalline specimen, 3 1 - M2 = .5472”? ; I(I+1)E rjg 4 2 2 1 -6 * 13'an Ifufdkzrjf ' (166) where the first term and the second term are the contributions from the homonuclear and the heteronuclear dipolar interactions, respectively, and n is the number of equivalent nuclei in a unit cell. Since the indices j, k and f run over every nucleus in the lattice, it is more convenient to write the second moment as the sum of two terms, M = M 2 (167) 2,intra + M2,inter’ where the first term is the contribution from the nuclei within the molecule and can be readily calculated if the molecular geometry is known. The second term contains long-range interactions from nuclei at remote lattice sites and is more difficult to evaluate. Andrew266 has considered the intermolecular contribution from remote nuclei with rij>M as a continuous distribution of spins. M, the truncating radius, depends on the lattice structure and 5 X has been proposed as a proper 157 value. With this choice, the intermolecular contribution to the second moment of fluorine nuclei can be estimated by -6 4nN M . = 317.2 Z N. Z n. 2 r. + 317.2 -——, (168) 2,1nter i 1 j J k Jk 3M3V slh‘ where n is the number of fluorine nuclei in a molecule, Ni is the number of molecules on the i£h_nearest neighbor sites, nj is the number of jth_fluorine nuclei in the origin molecule which are equivalent to an ith_neighbor, rjk is the distance between fluorine j of the origin molecule and fluorine k on the nearest neighbor, N is the number of fluorine nuclei per unit cell and V is the volume of the unit cell in cubic Angstrons. When the molecules in a lattice undergo an isotrOpic reorientation- al motion, the intramolecular part of the second moment vanishes and the intermolecular part can be simplified to = 317.2 n2 Niric’, (169) 1 M2,inter where E Nir;6 has been calculated to be 29.045 a.6 and 115.631 a.6 ( a is the lattice constant) for body-centered cubic and face-centered cubic crystal structures, respectively. When self-diffusion occurs in the plastic phase, this intermolecular contribution to the second moment will be averaged out and the linewidth will approach that in the liquid phase. The effect of lattice vibration on the NMR second moment has 267,268 been discussed by Shmueli, Polak and Sheinblatt . From their calculations and comparison with the experimental results for a few 158 compounds, it was found that these vibrations of the lattice tend to decrease the intramolecular part, but increase the intermoleculat part, of the rigid lattice second moment and the total effect does not exceed 12%. 3. Spin-Lattice Relaxation Times Since the linewidth is only sensitive to those molecular motions with correlation frequencies of the order of the linewidth in the solid phase, i.e., 103 to 105 Hz, only slow motions of the molecules can be studied by linewidth measurement. Spin-lattice relaxation times, on the other hand, have a much shorter time scale, namely, the reciprocal of the Larmor frequency, and are therefore more sensitive to faster molecular motions. Bloembergen, Purcell and Pound52 have derived a general equation for homonuclear dipolar relaxation in terms of the power spectra of the nuclear position functions which are random. functions of time: 3 4 2 1 R1 = 27 h I(I+l) J1(w°)+ 552(2wo) , (170) where the spectral .densitiesTJ1(u) are the Fourier transform of the correlation functions 1-3cos20..(t) Go(t) = < X | 3 1’ }> (171) j rij(t) sinze..(t)ex (21¢..(t)) Gl(t) = < Z 13 p *1. !> (172) . 3 J rij(t) and 159 sine . . (t) cose . . (t) exp(-i¢ . . (t)) 62(t) a < Z | ' 13 11 13 i) J 3 , (173) rij (t) in which r, 0, and 0 are the spherical polar coordinates of spin j relative to spin i. If the molecules undergo a rotational Brownian motion, the internuclear distance is fixed and the correlation functions can be expressed by exponential functions with a time constant Tc, which is the reorientational correlation time. Therefore Equation (170) becomes T 4Tc R1 = 1% Y4MZI(I+1)( (2: 2 + j?) i—Z Z ri?’ (174) l+morc 1+4moTc i jfii where indices 1 and j run over all equivalent nuclei in the lattice. By combining this equation with the second moment expression, a more convenient equation can be derived, that is, R --3 2M ( 1° 4 41° ) (175) l ' 3 7 2r 1+ 2 2 1+4 2 2 ' “oTc onc where M21. is the second moment modulated by the molecular motion considered. When the molecules undergo a translational motion,rij will also 53’269 has considered a random- be a random function of time. Torrey flight model for the self-diffusion in a crystal lattice and derived an equation for the spin-lattice relaxation rate as follows: = 8fly4H2I(I+1)nc Skslsw R 9(k.y). (176) 1 where n is the number of molecules per unit volume, c the ratio of the number of nuclei to the number of lattice sites, k the ratio of the 160 closest possible distance a of two approaching nuclei to the length of each diffusion jump t; y is equal to %wtd and rd, the mean time between adjacent diffusion jumps, is related to the self-diffusion coefficient Ds by Equation (61). w(k,y) has been calculated for various values of y and for both bcc and fcc lattices. Equation (176) can also be written in terms of the second moment as follows: 2 R, = v 112de {C(k.y)+4cck.2y)}, (177) where M2d is the second moment modulated by the self-diffusion,and the 6 functions are calculated and tabulated by Torrey53’269. It should be noted that by writing the spin-lattice relaxation rate in the form of Equation (176) or (177), the contributions from both the homonuclear and the heteronuclear interactions have been considered. 4. Spin-Lattice Relaxation Times in the Rotating Frame _ In deriving the Bloembergen-Purcell-Pound equations, perturbation theory was used since the fluctuating dipolar field is much weaker than the static field (~10 kG) applied. When the static field is low enough that the dipolar interaction becomes comparable to it, the perturbation treatment of Bloembergen-Purcell-Pound is no longer valid. Slichter and Ailion27o'273 first considered this strong-collision case and devised an experimental technique, namely, adiabatic de- magnetization in the rotating frame271 , to measure the spin-lattice relaxation time in the presence of a moderately strong field H1 after the static field Ho has been pulsed off resonance. The measured 161 relaxation rate Rlp was found to be Hi+aH§ H; R = R'(—-—) + R (——). (173) 10 1 2 2 c 2 2 H1+HD H1+HD where HD is the local dipolar field strength, Ri is the spin-lattice relaxation rate due to all mechanisms other than molecular jumping in the lattice, a is a number which determines the effect of the spin- lattice relaxation on the dipolar relaxation and Rc is the dipolar relaxation rate due to the molecular jumping. The technique of adiabatic demagnetization in the rotating frame was found to be inadequate for measuring short T1p values. An alterna- tive method has been developed by Hartmann and Hahn274. In this method, the alignment of the magnetization along H1 is obtained by a 900 pulse and followed immediately by a 900 phase-shifted H1 pulse with variable duration in order to measure the decaying magnetization at different times. Several other methods have also been developed, each with its own advantages and disadvantage5275. When the applied spin-locking pulse field H1 is much smaller than the local dipolar field HD, perturbation theory can still be used. 276 277 Look and Lowe , and also Jones , have derived an equation for R1 in this weak-collision case. For molecules undergoing an isotropic random motion with a correlation time T Rlp can be written as c! 5 T T 3 1’ _. 22.—:2... c ), (179) 2 l+motc 1+4wotc 2 O R19 3 K( where K is a nuclear constant and w1=yH1. In the case of Tcwo>>1, Equation (179) becomes 162 3 Tc R19 3K :77 . (180) m17c From the above equation, it can be seen that R1p is sensitive to molecular motions with a time scale of the order of 1/m1. By adjusting the spin-locking field strength H1, a wide range of Tc values can be determined. Accordingly, this should be very useful in studying the self-diffusion of the molecules in plastic crystals. Using Torrey's lattice-diffusion m0d8153’269, 245 R10 of a nucleus in a plastic solid phase was found to be - §_ Td 5 rd R10 " 7.2"2de1 2 50921013143 60,603) +G(k,zuo%9)}, (181) where the symbols are defined as before. When the diffusion is slow, such that mon>>1, Equation (181) becomes R1p . g-szzdrd 6(k,261;9) (182) and R1p has a maximum value when wltd=0.86. C. NMR Studies of Some Fluorochloroethanes ngnglé-n Due to their inert chemical properties, some of the fluorochloro- ethanes have been widely used as refrigerants and heat transfer fluids as well as for other industrial purposes. For many years, NMR workers have been interested in studies of the internal rotation of the molecules 163 279’291. However, very few people have tried to investi- in these Freons gate the overall molecular motions in both solid and liquid phases of these compounds, many of which form an interesting plastic phase in the solid state. In Table 30, we list some physical properties of these compounds. It can be seen that C2F6, CF3CC13, CFZClCCls, CFClZCFClz and CFClZCCls all have a very small liquid range (less than 55°C) which is one of the characteristic prOperties of plastic crystalline substances. In the following sections, we report our studies of molecular motions in some of these fluorochloroethanes by means of 19F NMR lineshape analysis and relaxation rate measurements. 1. Molecular Motions in CFZCICCl3 19 a. Measurements of the P NMR Linewidths and Second Moments The linewidths and second moments of the 19 F absorption in CFZClCCl3 at different temperatures are listed in Table 31 and plotted in Figure 29. A solid-solid phase transition at ~-115°C is very obvious since both the linewidth and the second moment drop rapidly at that temperature. After this transition point, the second moment reaches a plateau with a value of ~0.15 62. At temperatures higher than -60°C, both the linewidth and the second moment decrease slowly to zero. This is evidently due to the increasing self-diffusive motion which further averages out the intermolecular dipolar interaction. Experi- mentally, we found that the lineshape was changing from a Gaussian to a Lorentzian shape as the temperature increased in this region and the second moment became zero far below the melting point. 164 a Table 30. Physical constants of some fluorochlorocarbons. Compound Transition Meltigg Boiligg Critical point point point point *cpscp3 -101 -78 33.3 CF3CF2C1 -192.8 -106 -38 80.1 CF3CFC12 -56.6 3.6 CFZCICFZCI -94 3.6 145.7 *CF3c013 -118° 14.9 45.9 210c CFZCICFCIZ -35 47.6 214 *CFClZCFClz 26.5 91 278 *cpzmca3 -115d 40.6 91 *CFCIZCC13 100 136.8 * CF3C12CCC13 109.1 153.1 *CF3C12CCC12CF3 83 131 d * - CFZCCIZCClZCFZ 70 84.8 131.6 aAll temperatures are in °C. bReference 278. cReference 131. dThis work. *Plastic crystal. 165 Table 31. 19F linewidths and second moments in CFZClCClS. T 1000/T Ava M2 (°C) (°x 1 (Hz) (62) -43.5 4.36 1465 0.040 -51.0 4.50 3051 0.141 -56.0 4.61 4040 0.141 -62.0 4.74 4330 0.164 -68.0 4.88 4444 0.153 -73.0 5.00 4786 0.142 -78.0 5.13 4712 0.148 -86.0 5.35 4932 0.158 -94.0 5.59 5249 0.150 -100.0 5.78 5128 0.147 -108.0 6.06 5200 0.184 -llS.0 6.33 5670 0.207 -121.0 6.58 7790 0.592 -127.0 6.85 7373 .0.960 -116.0 6.37 8300 0.875 -llS.0 6.33 5740 0.785 -ll3.0 6.25 5029 0.157 -34.0 4.18 635 0.027 -33.0 4.17 635 0.013 -18.0 3.92 125 0.001 aAv(Hz) = 4007.4 X Av(G). 166 2 O o O A V —0.9 0 M2 10.8 1.5— -0.7 O O -0.6 1:? .2." U) U) a 8 3| — ~41-0.5 3 N <1 2 0.4 D 0.3 0.51— —0.2 ‘001 l I l -120 -100 -80 -60 t.°c 19 -4O Figure 29. F NMR linewidths (scale at left) and second moments (scale at right) of CFZClCClS. 167 From the linewidth narrowing in the temperature region from -70°C to ~0°C, the mean jump time Id can be obtained by using Equation (162) or (163) or (164). These results are plotted in Figure 32 and will be discussed later. The second moment observed in the plateau region (see Figure 29) is the contribution from the intermolecular dipolar interaction,which can also be calculated theoretically by using Equation (169). Unfortunately, no crystal structure has been reported for this compound, therefore, we can only estimate the lattice constant from the density by assuming either the body-centered or face-centered cubic structure for the crystal. We have measured the density d of this compound at room temperature and found it to be ~1.719 g cm's, hence, the lattice constant a will be 37.33 X if the crystal structure is body-centered cubic and ~9.24 X if it is face-centered cubic. Therefore, the second moment M2d due to the intermolecular dipolar interaction of 19F nuclei can be calculated to be @0.119 G2 (bcc) or ~0.118 G2 (fcc). The contribution from chlorine nuclei can be estimated to be about 4.5% of the above values, hence the total intermolecular second moment is ~0.124 62 which is in good agreement with the experimental result. b. Rotational Motion in the Plastic Phase The spin-lattice relaxation rates have been measured over a wide range of temperatures. These are listed in Table 32 and plotted in Figure 30. A discontinuity in the measured spin-lattice relaxation rates signals a solid-solid phase transition which occurs at N-IISOC; this agrees with similar discontinuities observed in the linewidth Table 32. 19 168 F spin-lattice relaxation rates in CF2C1CC13. T 1000/T R1 (°C) (°x'1) (sec-1) -60.0 4.69 0.1087:0.0023 -65.0 4.81 0.134410.0034 -79.0 5.15 0.1518:0.0014 -73.5 5.01 0.125810.0048 -73.0 5.00 0.125710.0036 -69.0 4.90 0.125410.0044 -67.0 4.85 0.1320:0.0040 -62.0 4.74 0.106010.0030 -77.0 5.10 0.1560i0.0050 -86.0 5.35 0.1850:0.0030 -58.0 4.65 0.102010.0020 ~50.0 4.48 0.0938:0.0022 -42.0 4.33 0.0933:0.0030 -33.0 4.17 0.0752:0.0018 -22.0 3.98 0.0778:0.0011 -43.0 4.35 0.0913:0.001S -l3.0 3.85 0.0861:0.0010 -2.0 3.69 0.095510.0015 6.0 3.58 0.0922:0.0018 -2.5 3.70 0.0861:0.0014 16.0 3.46 0.094410.0036 35.0 3.25 0.1300:0.0060 41.0 3.18 0.1340:0.0070 -6.0 3.74 0.078110.0016 -11.5 3.82 0.0805:0.0024 -18.S 3.93 0.0860:0.0021 -25.0 4.03 0.0900:0.0028 -31.0 4.13 0.0935:0.0021 -32.0 4.15 0.0863:0.0012 -38.0 4.26 0.088510.0020 -43.0 4.35 0.0945:0.0027 -43.5 4.36 0.0917:0.0016 -51.0 4.50 0.103210.0048 -56.0 4.61 0.108110.0112 -62.0 4.74 0.1300:0.0046 ~68.0 4.88 0.145010.0070 -73.0 5.00 0.167010.0063 -78.0 5.13 0.1930:0.0090 -86.0 5.35 0.2610:0.0090 -94.0 5.59 0.5720:0.0270 Table 32. (cont'd) 169 T IOOO/T R1 (°C) (°x'1) (sec‘l) -100.0 5.78 0.4150i0.0110 -108.0 6.06 0.5720:0.0270 -115.0 6.33 0.7700:0.0570 -121.0 6.58 0.1680i0.0170 -127.0 6.85 0.129010.0040 -136.0 7.30 0.0518:0.0040 -144.0 7.75 0.0203:0.0037 -116.0 6.37 0.2770i0.0090 -120.0 6.54 0.1980i0.0120 -117.0 6.41 0.1560:0.0110 -117.5 6.43 0.1670i0.0140 -115.0 6.33 0.1530:0.0110 -113.0 6.25 0.6400:0.0120 -12.0 3.83 0.0868i0.0021 23.0 3.38 0.094910.003S 29.0 3.31 0.1153i0.0031 17.0 3.45 0.0978i0.0022 11.0 3.52 0.0986:0.001S 4.0 3.61 0.0894i0.0012 -3.0 3.70 0.0882:0.0016 -21.0 3.97 0.0894:0.0030 -22.0 3.98 0.0849i0.0012 63.0 2.98 0.1480:0.0083 80.0 2.83 0.1560i0.0090 103.0 2.66 0.1820:0.0130 132.0 2.47 0.224010.0220 170.0 2.26 0.2850:0.0170 207.0 2.08 0.380010.0290 170 \ t . 0 1—‘., ,. \ 9. .. -\ . : 2, i 05- 3 ‘\ r v—A 1.. q? n?‘p E v 9. . .. | I GE I 1 0-2C . ' - 1 '\ \. '. I . 1 0‘ \. . ' . e \ \ . ' \ \. ',' 0 ~ * \I . \ 0.1 P '\. e . . . . . \\ \ I. \. 0. e . \- \ .\ 005- x \e .\O \0 \e \e \e 002% \ \ l l J L l 2 3 4 5 6 7 IOOO/T Figure 30. Plot of the 19F relaxationrates in CFZClCClg as a functirim of reciprogal temperature. 0 : R}; 0: R5; ---. are. _ e -0- 51'- one. SC . R1 , . R1, . R1 , . 32 171 and the second moment measurements. A few R1 values obtained at temperatures below the transition point imply the existence of a molecular motion with an activation energy about 4.1:0.2 kcal/mole. This is probably the hindered internal rotation of the CFZCl group of the molecule with respect to the pseudo C3 symmetry axis along the carbon-carbon chemical bond. To describe this motion, we consider only the intramolecular dipolar interaction between two fluorine nuclei separated by a distance r¢2.l6 2 (assuming rCF-z 1.33 X and _/_ FCF=108° as in szcmpzazgz). The spin-lattice relaxation rate due to this slow motion, which has morc>>1, can be derived from Equation (174) and is found to be 1.r. _ 3 4 2 -6 1 ~ -8 R1 - 'S—Y n 1' T - 2.09x10 /Tc. (183) “01c Therefore, the correlation time T of this internal rotation is C calculated to be Tc 2 1.16XI0-13exp(4.1 kcal mole'I/RT). (184) In the temperature range between -115°C and -63°C, the spin-lattice relaxation rates show a fairly linear dependence on l/T and have an activation energy ~2.S6:0.10 kcal/mole. The motion which produces the largest contribution to the relaxation is the overall tumbling motion of the molecules with a correlation time Tr short snough so that 'onr<<1. From Equation (175), the spin-lattice relaxation rate in this temperature region can be written as 172 (185) where M21. is the dipolar second moment averaged out by this overall tumbling motion and can be estimated from Equation (166) by considering only the intramolecular contribution. We obtained M2r=3.1 G2 and Rr ~ 6 55 109 186 1-... .r. (1 Therefore, the correlation time T1. is calculated to be 1 T 2 3.6QXIO- 4exp(2.56 kcal mole-I/RT), (187) r which gives rr=2.25 psec at the melting point. As the temperature increases to higher than -63°C, the relaxation rates start to deviate from the linear plot of the intramolecular relaxation rates versus reciprocal temperature. This is a result of contributions from additional relaxation mechanisms such as the self- diffusion of the molecules and the spin-rotational interaction. In order to separate these two contributions, we first try to estimate the effect of self-diffusion. Assuming that the lattice-diffusion model is valid and that wOTd>>1, so that G(k,y)+4G(k,2y)=0.5618/y269, the spin- lattice relaxation rate due to the self-diffusion can be calculated from Equation (177) to be 8 M26 ‘td' R: = 1.12 72M -—————-= 1.13x10' word/2 (188) Using the measured M2d=0.15 62, Equation (188) becomes 173 Rf = 1.69><10'g l—. (189) T6 We will discuss this later after the values of Td are obtained. c. Self-Diffusion in the Plastic Phase Since the spin-lattice relaxation in the rotating frame has the right time scale to describe the self-diffusive motion in plastic crystals, we measured Rlp values at three different spin-locking fields (22.5 G, 5.7 G and 4.5 G). These are shown in Table 33 and are plotted in Figure 31. The maxima of the curves in Figure 31 occur when wle 20.816. Using Equation (182), the second moment modulated by the translational motion can be obtained from these maxima. This is found to 2 2 at the three different have the values 0.154 62, 0.151 G and 0.156 G spin-locking fields used. The agreement of these values with the measured one ( 0.15 62) is very good. Assuming that the molecules undergo a self-diffusion via vacancy migration, the mean jump time Td can be derived from the measured R1 values using Equation (182). These are plotted versus reciprocal temperature in Figure 32 together with those derived from the linewidths, and from the spin-spin relaxation rates,listed in Table 34. As can be seen from this plot, the mean jump times derived from the linewidths depend on whether Equation (162), Equation (163) or Equation (164) is used and can differ by a factor of two to ten. Also, the low temperature values deviate from linearity. Therefore,reliable information about the self-diffusive motion cannot be obtained solely from linewidth measurements. The best least-squares fit to these Td values can be 174 Table 33. 19F spin-lattice relaxation rates in the rotating frame in CFZClCClS. T 1000/T R1p ug (°C) (°x'1) (sec‘l) (kHz) 29.0 3.31 5.5410.11 90.0 15.0 3,47 12.98i0.15 90.0 6.0 3.58 22.37i0.05 90.0 -2.0 3.69 32.8910.49 90.0 -10.0 3.80 36.77iO.30 90.0 -17.0 3.91 28.5010.41 90.0 -23.0 4.00 20.54i0.39 90.0 -30.0 4.12 13.55i0.34 90.0 -36.0 4.22 9.820i0.23 90.0 -42.0 4.33 6.160i0.24 90.0 28.0 3.32 6.820t0.20 27.0 17.0 3.45 11.36i0.33 27.0 8.0 3.56 24.92i0.49 27.0 -1.0 3.68 52.77to.51 27.0 -9.0 3.79 86.4811.35 27.0 -16.0 3.89 114.9:2.5 27.0 -23.5 4.01 114.1i1.8 27.0 -30.0 4.12 92.10i1.13 27.0 -36.0 4.22 61.00:].03 27.0 -43.0 4.35 40.51i1.27 27.0 -53.0 4.54 . 21.87i0.77 27.0 20.0 3.41 11.05:0.58 18.0 11.0 3.52 22.16:0.28 18.0 1.0 3.65 47.0810.89 18.0 -7.0 3.76 91.96il.29 18.0 -15.0 3.88 154.2:2.6 18.0 -22.0 3.98 188.1:3.5 18.0 -29.0 4.10 180.314.1 18.0 -37.0 4.24 125.212.2 18.0 -43.0 4.35 74.51il.49 18.0 -S0.0 4.48 40.41:0.93 18.0 -55.5 4.60 33.0211.57 18.0 -60.0 4.69 20.6811.26 18.0 8H2(kHz) = 4007.4XH2(G). 175 Table 34. 19F spin-spin relaxation rates in CF2C1CC13. T IOOO/T R2 (°C) (°K“) (sec'l) 207.0 2.08 2.89:0.23 170.0 2.26 0.788:0.046 132.0 2.47 0.526t0.041 103.0 2.66 0.410i0.026 80.0 2.83 0.37810.023 63.0 2.98 0.301i0.022 49.5 3.10 0.270i0.006 34.0 3.26 4.18:0.26 26.0 3.34 10.36i0.98 12.0 3.51 45.5:2.8 1.0 3.65 133:5 -6.0 3.74 372:29 -18.0 3.92 393130 -33.0 4.17 19951210 -43.5 4.36 46021418 176 I‘D‘N3i \ \ o’ '6. 3 Im '" .30 \\ o, \x . \ \ \ A o G \ To . \ a: ' Q a, g \ 4.5 G. ‘1; \. '. \ (f b, 9 b .\ .2 \\ \ '. \ 9,. 6.76 I o —- ‘y K. ‘. 225 G. '\ ' L 1 3 4 5 IOOO/ T Figure 31. Plot of the 19F spin-lattice relaxation rate in the rotating frame at three different spin-locking fields,and the spin-spin relaxation rate (———),in CF2C1CC13. 177 o o o 16 - j o 16‘— 4 53 .3 .5 . IO t- .1 i 10‘- 4 -7 '0 H -< 1 3 IOOO/T Figure 32. Plot of the mean jump time for self-diffusion in the plastic phase of CFZClCCl3; e : from R19 data; 0: from the linewidth data using Equation (162); o: from the linewidth data using Equation (163); 0: from the line- width data using Equation (164); +: from R2 data. 178 written as -16 -1 Td = 5.8XI0 exP(11.6 kcal mole /RT), (190) which gives -rd=*6.7><10'8 sec at the melting point. For a monovacancy diffusion mechanism, the mean jump distance is the distance to the . . . /' . closest lattice Sites and 15 equal to —%-a for a body-centered cubic structure and 2%; for a face-centered cubic structure. Using Equation 2 (61), the self-diffusion coefficient at the melting point is ~1.OXI0-8 cm2 sec'1 or '\al.06><10-8 cmzsec'l, depending on whether the crystal structure is body-centered cubic or face-centered cubic. With the Td values expressed by Equation (190), we can estimate the contribution to the spin-lattice relaxation rate from the inter- molecular dipolar interaction. This can be readily done by combining Equation (189) and Equation (190) to give 5 6 1 ‘1 R1 = 2.QZXIO exp(-11.6 kcal mole /RT). (191) The calculated values of R: are about 17% of the total spin-lattice relaxation rate near the melting point and become negligible at temperatures below 0°C. 19 3 d. F Relaxation in the Liquid Phase and J(F sC1) In the liquid phase of CF2C1CC13 the molecular motions are expected to be very isotropic in view of the near-spherical molecular shape (the three principal components of the moment of inertia are 179 "4°g cm2). From Figure 30, it 74SXI0'40g cmz, 851x10'4°g cm2 and 957X10 can be seen that the spin-rotational interaction dominates both the spin-lattice and the spin-spin relaxation. The difference between the spin-spin and the spin-lattice relaxation rates is due to different contributions from the scalar-coupling interaction, which has a negligible effect in determining the spin-lattice relaxation rate but is important for the spin-spin relaxation rate. Therefore, we can write 3 2 3S 2 37 37 35 Rz-Rl = age = 10n2( %-J (F Cl)T1( Cl)+z-J (F Cl)T1( 01)) = 95.7 J2(F35C1)T1(°5Cl), (192) where we have used Equation (70) for each isotope and relate the coupling constants and the spin-lattice relaxation times of two chlorine isotopes by their gyromagnetic ratios and their nuclear quadrupole moments, respectively. The chlorine spin-lattice relaxa- tion time can be calculated from the quadrupole coupling constant and the reorientational correlation time of the molecule as described by Equation (67). Although the quadrupole coupling constant of 35C1 in this compound has not been reported, a value of 78 MHz, which lies between those found in CFZCl2 and CF3C1, should be an adequate estimate since the chlorine quadrupole coupling constants in other chloromethanes and chloroethanes do not vary appreciably219. We assume that the reorientational correlation time can be estimated from Equation (187), since for plastic crystals it is very likely that the reorientational motion of the molecules in both liquid and 180 plastic solid phases follows the same mechanism. Therefore, we obtain -1 35 -1 T1 ( C1) = 780 exp(2.56 kcal mole /RT) (193) and the coupling constant J(F35C1) can be calculated from Equation (192) and Equation (193) giving J(F35C1)28.1 Hz. Since the necessary physical constants have not been reported for this compound, the contribution of the intermolecular dipolar inter- 19 action to the F relaxation in the liquid phase was not evaluated. However, as in CFSCCIS, which has been previously studied131, this contribution should be minor but just enough to smooth out the relaxation rate plot shown in Figure 30. The significance behind this is that the reorientational motions in both liquid and plastic solid phases are via the same mechanism. The resultant relaxation rates, after subtracting the intermolecular contributions show two com- peting relaxation contributions with opposite temperature dependence. These are obviously those from the spin-rotational and the intra- molecular dipolar interactions. The spin-rotational relaxation rates can be readily separated since the intramolecular dipolar relaxation rates are available from Equations(186) and (187). Assuming that the molecules are spherical-top rotors, the spin-rotational relaxation rates can be expressed by Equation (80). We have calculated the average moment of inertia to be 8SOX10‘4°g cm2 and the effective spin-rotation constant Ceff to be ~1.17 kHz from the 19F chemical shielding (66.8 ppm downfield from CFC13). Therefore, sr R1 R 1.14X109TTJ (194) 181 HewHV :oHuascm eH Hemoe a .HemHU :eHuesem segue MN0.0 0.vN vw.H H.vv mv0.0 000.0 m.m NmH wmo.0 N.MH mh.H m.MN 000.0 mH0.0 0.m 00m 0m0.0 mn.h 00.H N.NH 000.0 mN0.0 m.v NNN 000.0 00.v hm.H mm.0 HmH.0 mv0.0 0.¢ 0mm mH.0 0N.N hv.n mm.n vNN.0 n50.0 m.n 0mm vm.0 0N.H 0m.H 0b.H mm.0 mNH.0 0.n nnm Hm.0 mm.0 mm.~ 0N.H 0v.0 BH.0 h.N 05m wm.0 «5.0 QN.~ Nm.0 he.0 mHN.0 m.N 00v «v.0 50.0 NN.H Hw.0 0m.0 v~.0 v.N hfiv 00.0 00.0 0H.H Hh.0 nm.0 m0N.0 n.N mmv Nm.0 vm.0 0H.“ m0.0 00.0 Hm.0 N.N vmv 00.0 wv.0 QH.H mm.0 00.0 hn.0 H.N 059 05.0 vv.0 HH.H 0v.0 vw.0 00.0 0.N 00m noommv noommv noommv nHtoomv AHuxoV axov me me xHex\HV Me. eee ewe e\oooH a .nHuu Hu~au eH meaHe eoHumHouuou HecoHuapcoHnoou on» use emueoeoa ueHmmee 0gp mo :oHumHmoHeu on» we Aheaamm .mm oHnmh 182 Figure 33. Plot of reduced reorientational correlation time versus reduced angular momentum correlation time for CFZClCCl3; solid lines A and C represent the theoretical curves predicted from the M-diffusion model for i=1 and 2,res- pectively; solid lines B and D represent the theoretical curves predicted from the J—diffusion model for i=1 and 2, respectively; dashed lines represent the theoretical result predicted from the Debye classical diffusion model (TeTJ=I/2(2+l)kT); dotted lines represent the theoretical result predicted from the perturbed free-rotor model (T9=TJ/22+l) and the broken lines represent the free-rotor limit of the extended diffusion model (T9=TJ/2£). 183 0.0 t «.0 H0 00.0 184 and the effective angular momentum correlation time TJ can be estimated and is shown in Table 35. In Figure 33 we plot the reduced reorientational correlation times T5 versus the reduced angular momentum correlation times 13. It is obvious that the experimental data points do not follow the J-diffusion or the classical diffusion model. 2. Molecular Motions in CFCIZCFCIZ a. Measurement of the 19F NMR Linewidths and Second Moments CFClzCFClz is an isomer of CFZCICClS. Both compounds have the same boiling point but the melting point of CFC12CFC12 is lower, which gives a wider liquid range and therefore a smaller plasticity is 9F NMR linewidths and the second moments are expected. The measured 1 listed in Table 36 and are both plotted versus reciprocal temperature in Figure 34. No obvious phase transition was observed above -130°C. The intermolecular dipolar second moment was found to be ~0.18 G2 which decreased to zero gradually at temperatures above -90°C. The theoretical estimate of this intermolecular dipolar second moment was ~0.lZ7 Gz( if the crystal has a body-centered cubic structure) or ~0.125 G2( if the crystal has a face-centered cubic structure). These values are obtained using the measured solid densityof 1.735 g cm"3 at 5°C, which givesa lattice constant of 7.31 X or 9.21 X,depending on whether the crystal structure is body-centered cubic or face-centered cubic. 185 Table 36. 19F linewidths and second moments in CFClZCFClz. T 1000/T 00' M2 (°C) (°x'1 (Hz) (62) 23.0 3.38 14.6 0 7.0 3.57 26.0 0 0.0 3.66 38.0 0 10.0 3.53 24.0 0 -2.0 3.69 41.0 0 -10.0 3.80 60.0 0 -ll.0 3.82 63.0 0.0003 -22.0 3.98 107.0 0.0008 -23.0 4.00 102.0 0.0009 -32.0 4.15 224.0 0.0027 -33.0 4.17 229.0 0.0026 -39.0 4.27 395.0 0.0048 -40.0 4.29 419.0 0.0062 -43.0 4.35 585.0 0.013 -46.0 4.40 779.0 0.015 -53.0 4.54 1374.0 0.053 -S8.0 4.65 2515.0 0.100 -64.0 4.78 3370 0.126 -72.0 4.98 4035 0.155 -77.0 5.10 4300 0.165 -83.0 5.26 4540 0.180 -90.0 5.46 4675 0.180 -96.0 5.65 4690 0.180 -106.0 5.99 5020 0.200 -107.0 6.02 4950 0.190 -llS.0 6.33 4830 0.180 -122.0 6.62 4950 0.190 186 0 AI' (9 0 M2 I .- -—0.2 3? .6" 2 £5 < a 43 Q N :5 0.5" “'10.! -100 -50 o TEMP.(°C) Figure 34. 19F NMR linewidth (scale at left) and second moment (scale at right) for CFClZCFClz. 187 b. Rotational Motion in the Plastic Solid Phase The measured 19F spin-lattice relaxation rates are listed in Table 37 and plotted versus reciprocal temperature in Figure 35. At low temperatures the plot shows fairly good linearity with a slope ~3.l6 kcal/mole. The most probable interactions contributing to relaxation in this temperature range are the intramolecular dipolar interaction and the interaction through the chemical shielding aniso- tropy. The intramolecular dipolar relaxation rate can be expressed by Equation (175). The intramolecular dipolar second moment M2r can be claculated from the geometry of the molecule by assuming a transfform, which was found to be ~16515 cal/mole more stable than the gauche-form in CFC13 solution287. We calculate MZr to be ~0.23 G2 which gives d R 8 l = 4.SX10 Tr. (195) By assuming that the overall tumbling motion of the molecules is isotropic, we can use the same Tr to evaluate the relaxation rate due to the chemical shielding anisotropy A0; the latter has been measured in polycrystalline CFClZCFCl2 and found to be ~240 ppm293 (in the C-F bond axis system). This gives the relationship C58 8 1 = 9.6x10 Tr. (196) R Therefore, the low temperature linear part of the spin-lattice relaxation rate plot can be expressed by the sum of Equations (195) and (196), and the reorientational correlation time Tr can be determined to be 188 Table 37. 19F spin-lattice relaxation rates in CFClZCFClz. T 1000/T R1 (°C) (°K'1) (sec‘l) -132.0 7.09 0.700:0.123 -13l.0 7.04 0.51310.089 -128.0 6.90 0.566:0.082 -122.0 6.62 0.412:0.042 -121.0 6.58 0.30210.024 -114.0 6.29 0.276:0.015 -113.0 6.25 0.180:0.018 -104.0 5.92 0.122:0.010 -103.0 5.88 0.100:0.017 -97.0 5.68 0.075:0.004 -88.5 5.42 (0.07910.006 -80.5 5.19 0.05410.006 -73.0 5.00 0.061910.0033 -63.0 4.76 0.0457:0.0012 -S7.0 4.63 0.0503:0.0035 -37.0 4.24 0.0473:0.0010 -14.0 3.86 0.0510:0.0010 2.0 3.64 0.0622:0.0025 23.0 3.38 0.0931:0.0015 25.5 3.35 0.097410.0060 41.0 3.18 0.1028:0.0020 44.0 3.15 0.1020:0.0090 57.0 3.03 0.1295:0.0059 75.0 2.87 0.1397:0.0010 102.0 2.67 0.155:0.001 129.0 2.49 0.18210.002 149.0 2.37 0.21210.005 189 \Q .90 ME I" 1&3) I .\c; '3. I ‘- '\ | °\. 45" : ‘-\‘ I. I ‘4‘ L) I . m . U° I \_z ' (I , / '/ OJ " 0 I I II I I I I I I l 2 3 4 . 5 6 7 IOOO/T Figure 35. 19F spin-lattice relaxation rate (0) and the spin-spin relaxation rate (0) in CFCIZCFCIZ; the straight solid line represents the intramolecular dipolar relaxation rate R9; --- : the scalar spin-spin relaxation rate Rgc. 190 T = 5.8x10'l r Sexp(3.l6 kcal mole'I/RT). (197) This equation gives Tr=1.21 psec at a temperature just below the melting point. c. Self-Diffusion in the Plastic Phase From the gradual decrease of the linewidth, we expect that the self-diffusive motion of the molecule has reached a rate of the order of the linewidth,~5000 Hz at temperatures higher than -100°C. We have measured the spin-lattice relaxation rates in the rotating frame at three different spin-locking fields, and also the spin-spin relaxation rates. These are listed in Tables 38 and 39 and are plotted versus reciprocal temperature in Figure 36. The mean jump times for the self-diffusive motion can be obtained from R19, R2 and Av as before. These are plotted in Figure 37 and show an activation energy ~11.4 kcal/mole. The best least-squares fit of these experimental data points can be expressed by Td = 1.9XI0-16exp(11.4 kcal mole-I/RT), (198) 8 which gives Td=4.SXl0' sec at the melting temperature. Again, by assuming a monovacancy diffusion mechanism, the self-diffusion 8 1 coefficient at the melting temperature is ~1.48X10' cmzsec' or 1.57x10'8 cmzsec'1 depending on whether the crystal structure is body-centered cubic or face-centered cubic. The maximum values of R10 give estimated intermolecular dipolar 191 19 , , Table 38. F spin-lattice relaxation rates in the rotating frame in CFCIZCFC12. T 1000/T R1 H2 (°C) (°x‘1) (sec'l) (kHz) 13.0 3.50 ll.55:2.28 41.5 -1.0 3.68 26.07:0.37 41.5 -10.0 3.80 55.1510.60 41.5 -16.0 3.89 68.09:1.07 41.5 -22.0 3.98 93.14:0.78 41.5 -28.0 4.08 93.77:1.04 41.5 -34.0 4.18 77.52:0.78 41.5 -40.0 4.29 52.60:1.02 41.5 -47.0 4.42 35.65:0.88 41.5 -54.0 4.57 22.48iO.69 41.5 -S4.5 4.58 23.65i1.19 41.5 -58.0 4.65 14.87i0.84 41.5 -59.0 4.67 16.69i0.55 41.5 -62.0 4.74 16.13:].07 41.5 -6S.0 4.81 11.81i0.58 41.5 -65.0 4.81 30.63il.O8 20.8 -60.0 4.69 46.74i0.24 20.8 -54.0 4.57 62.03iZ.38 20.8 -45.0 4.38 130.2:2.4 20.8 -39.0 4.27 190.1:5.8 20.8 -34.0 4.18 202.6:8.3 20.8 -27.0 4.06 159.3:2.4 20.8 -21.0 3.97 119.9i1.3 20.8 -13.0 3.85 71.7310.69 20.8 -1.5 3.68 31.05t0.48 20.8 23.0 3.38 4.2610.19 20.8 -10.0 3.80 54.88i0.86 8.2 -18.0 3.92 113.012.2 8.2 -26.0 4.05 236.6i5.0 8.2 -32.0 4.15 371.1:13.2 8.2 -39.0 4.27 _ 521.4i20.0 8.2 -4S.0 4.38 643.2:27.2 8.2 -51.0 4.50 437.0t21.6 8.2 -56.0 4.61 217.017.9 8.2 -60.0 4.69 157.4:10.8 8.2 -72.0 4.98 47.16i1.19 8.2 Table 39. 19 192 F spin-spin relaxation rates in CFCl 2 CFCIZ. T lOOO/T R2 (°C) (°K l1 (sec-1) -35.0 4.20 556.5i50.0 -23.0 4.00 208.1:20.8 -14.0 3.86 48.2i4.9 -5.0 3.73 34.7i1.9 1.0 3.65 17.2i0.8 7.0 3.57 13.1i0.6 11.0 3.52 8.810.3 17.0 3.45 6.1110.18 20.0 3.41 4.69i0.14 21.0 3.40 4.39:0.18 23.0 3.38 0.88:0.04 24.0 3.37 0.88:0.04 22.0 3.39 1.55i0.14 21.5 3.40 3.97:0.36 29.6 3.30 0.89610.053 49.2 3.10 1.00010.053 57.5 3.02 1.053i0.092 76.0 2.86 1.145i0.127 99.4 2.68 1.33910.101 122.0 2.53 1.542i0.122 193 Ioooh _ Ioo- ~ .2) DJ 3 0: IO — ~ I .. _ L 1 1 I 3.5 4 4.5 5 IOOO/T Figure 36. Plot of the 19F spin-lattice relaxation rate in the rotating frame at three spin-locking fields (10.5, 5.3 and 2.0 G) and the spin-spin relaxation rate (0),in the plastic solid phase of CFCIZCFClz. 194 A A A IC§4P A. . ‘ . . 1 .J g 1 1 6 IOOO/T Figure 37. The mean jump time derived from R10 (0), R2 (0) and Av (A) of 19F nuclei in the plastic solid phase of CFClZCFClz. 195 second moment of 0.18 62 , 0.19 G2 and 0.23 G2 which are in good to fair agreement with the directly measured value. The contribution to the spin-lattice relaxation rate from the translational motion in the plastic solid phase can also be evaluated from the second moment M2d and the mean jump time rd. The result is shown in Figure 35. 19 d. F Relaxation in the Liquid Phase and J(F35Cl) Accurate extraction of the spin-rotational relaxation rates was not attempted for this compound because the measurements were not carried to high enough temperatures. Nevertheless, several measurements of the spin-spin relaxation rate show that there is a large scalar relaxation contribution due to the fast-relaxing chlorine nuclei. This contribution should be about twice as big as that found in CFzClCC13 since there are two geminal chlorine nuclei coupled with each fluorine nucleus. By using the same approach as before, we obtain J(FSSCl) =11.7tl.7 Hz. 3. Self-Diffusion Studies in CFSCCIS CFSCC13 is the most symmetric compound in the fluorochloroethane series. The small liquid range (~32°C) and the relatively high melting point (14°C compared to -36°C for its asymmetric isomer CFZClCFClz) imply a great plasticity in its solid phase. In previous studies advantage has been taken of these properties to separate the possible contributions of different mechanisms to the 19F spin-lattice 31 relaxation in this compound1 . Also Gordon's extended J-diffusion 196 model was found to be appropriate to describe the reorientational motion of the molecules in both the liquid and the plastic solid phases. Furthermore, from Raman lineshape analysis it was found that the reorientational motion is nearly isotropiclsl. In order to investigate in more detail the self-diffusive motion in the plastic phase of this compound, we have carefully measured the linewidths and the second moments of the 19Fresonance lines and also the spin-lattice relaxation rates in the rotating frame. These are listed in Tables 40 and 41. In Figure 38 the linewidth and the second moment are plotted versus reciprocal temperature. We did not reach to 131 the transition point reported in the previous study because our digitizer (20 psec minimum dwell time) was not fast enough to follow the decay of the magnetization at low temperatures and to obtain enough data points for Fourier transformation. Nevertheless, a plateau was found in both the linewidth and the second moment plots at tempera- tures below 200°K. The second moment in this region was found to be 19 ~0.34 G2. Figure 39 shows the plots of the F spin-lattice relaxation rates in the rotating frame R at three different spin-locking fields 10 (17.6 G, 8.8 G and 5.7 G),and the 19F spin-spin relaxation rates. From the maxima of the R1 plots, the second moments due to the intermole- cular dipolar interaction were found to be 0.28 62, 0.29 G2 and 0.35 G2 at the three different spin-locking fields. The calculated value from the lattice constant of the crystal derived from the density (1.71 131 2 or 0.212 G€,depending on whether the crystal g cm'3) is ~0.213 G structure is body-centered cubic or face-centered cubic. The agreement with the experimental results is only fair. Using the same method as in previous sections, the mean jump time 197 Table 40. 19F linewidths and second moments in CF3CC13. T 1000/T Av M2 (°C) (°x‘1) (Hz) (02) -23.5 4.01 116 0.0033 -31.0 4.13 233 0.0058 -41.5 4.32 700 0.0185 -42.0 4.33 696 0.0232 -47.0 4.42 1155 0.0833 -50.0 4.48 1648 0.110 -55.0 4.59 2516 0.145 -59.0 4.67 3521 0.195 -64.0 4.78 4857 0.253 -68.0 4.88 5633 0.322 -74.0 5.02 6054 0.354 -78.0 5.13 6286 0.332 -84.0 5.29 6567 0.340 -88.0 5.40 6604 0.341 -94.0 5.59 6750 0.336 -101.0 5.81 6720 0.341 -111.0 6.17 6726 0.345 -120.0 6.53 6770 0.340 198 Table 41. 19F spin-lattice relaxation rates in the rotating frame in CFSCCls. T 1000/T Rlp H2 (°C) (°K'1) (seC’l) (kHZ) 4.0 3.61 10.70i0.20 35.5 -7.0 3.76 32.91i0.28 35.5 -17.0 3.91 79.2311.53 35.5 -28.0 4.08 163.Zil.7 35.5 -39.0 4.27 167.112.3 35.5 -48.0 4.44 75.44tl.44 35.5 -57.0 4.63 37.54:0.67 35.5 -64.0 4.78 21.36i0.37 35.5 6.0 3.58 10.3310.18 71.0 0.0 3.66 15.5410.17 71.0 -8.0 3.77 33.47:0.48 71.0 -16.0 3.68 55.1310.99 71.0 -23.0 4.00 86.28io.82 71.0 -3l.0 4.13 79.72:0.l7 71.0 -38.0 4.26 48.33tl.45 71.0 -44.0 4.37 31.6410.62 71.0 -50.0 4.48 18.96i0.52 71.0 -S6.0 4.61 10.0310.43 71.0 5.0 3.60 11.1910.15 22.7 -2.0 3.69 22.4110.14 22.7 -12.0 3.83 55.3810.94 22.7 -22.0 3.98 138.7il.3 22.7 -31.0 4.13 248.915.0 22.7 -39.0 4.27 335.3i2.1 22.7 -46.0 4.40 225.4i4.8 22.7 -53.0 4.54 129.7i3.0 22.7 -58.0 4.65 71.39i2.55 22.7 -62.0 4.74 53.54i2.34 22.7 -67.0 4.85 30.9310.65 22.7 -72.0 4.98 24.14i0.76 22.7 -76.0 5.08 14.71t0.l4 22.7 -81.0 5.21 12.26t0.4l 22.7 199 O 2 - O - 4.3 e) .g ‘9 . {I 0 LS- O A 8 . 2 . “-0.2 (D r-\ \_I (H e H 4 | +- 2 3 ’ N 2 I e -0.1 0.5 r- ' o A o 0 M2 0 / v, 7 I 1 4 5 6 IOOO/T Figure 38. Plot of the 19F NMR linewidth and second moment in CF3CC13 versus reciprocal temperature. 200 I"o“\ :0 3“ ,' a ""9. \ .. Q IOO- _, \ Ie»‘. ‘.K d} 25‘ '\ g f‘\ .1 .3 2°\ - \ G) \ ti 3. '3‘ 576. m \. .. \ . ‘\_l 3" ‘i? \ C]: \\ .'ELEI(i I o - ‘\ I7.6 G. i L, I - I 35 4.0 4.5 5.0 IOOO/ T Figure 39. Plots of the 19F spin-lattice relaxation rate in the rotating frame at three different spin-locking fields (17.6, 8.8 and 5.7 G), and of the spin-spin relaxation rate (0) versus reciprocal temperature, for CF3CC13. 201 '4 IO)- I U]— IOOO/T Figure 40. Mean jump time for self-diffusion in the plastic phase of CF3CC13 derived from R19 (0), R2(+) and Av (0) . 202 of the self-diffusive motion of the molecules in the plastic solid phase can be estimated from the spin-lattice relaxation rates in the rotating frame, the spin-spin relaxation rates and the linewidth narrowing. These are plotted in Figure 40. The best least squares fit to these experimental data points can be expressed by the equation -17 rd = 4.7x10 exp(11.7 kcal mole'l/RT), (199) which gives Td=4.0X10-8 sec at the melting point. The self-diffusion coefficient Ds at the melting point was found to be 1.61><10'8 cmzsec'1 or 1.70x10'8 cmzsec'i depending on whether the crystal has a body- centered cubic or a face-centered cubic structure. 4. Spin-Lattice Relaxation Studies in CFZClCF2§l_ The other two fluorochloroethanes, CF3CF2C1 and CFCIZCC13, also may form a plastic phase but data were not obtained in this work since the 19F relaxation in the former compound is complicated by the homo- nuclear coupling and the latter compound was not available for this work. We have also measured the spin-lattice relaxation rate of 19F in CFZCICFZCI, which is not expected to form a plastic solid phase. The results are listed in Table 42 and plotted versus reciprocal temperature in Figure 41. In the solid state, the low temperature R1 values are linear with reciprocal temperature and the activation energy is ~1.17 kcal/mole. Translational motion begins to contribute to the spin- lattice relaxation at a temperature about 18° below the melting point. 203 Table 42. 19F spin-lattice relaxation rates in CFZClCFZCl. T IOOO/T R1 (°C) (°x‘1) (sec‘l) -12.0 3.83 0.1803:0.003S -105.0 5.95 0.1643:0.0029 -111.0 6.71 0.1506:0.0027 -121.0 6.58 0.2013:0.0036 -128.0 6.90 0.230410.0050 -134.0 7.19 0.262810.0062 -135.0 7.25 0.256010.0115 -l41.0 7.58 0.3858:0.0217 -ll7.0 6.41 0.1797:0.0024 -103.0 5.88 0.1704:0.0050 -98.0 5.71 0.1881:0.0163 -9S.0 5.62 0.2175:0.0071 -91.0 5.49 0.234810.0088 -85.0 5.32 0.250410.0072 -78.0 5.13 0.185110.0024 -68.0 4.88 0.1694:0.0017 -60.0 4.69 0.192010.0088 -57.0 4.63 0.1603:0.0025 -40.0 4.29 0.1491:0.0042 -23.0 4.00 0.175610.0036 10.0 3.53 0.221210.0073 22.0 3.39 0.2969:0.0117 204 |_ .4 H C .. F‘ .. r d 1 M.P. . -. 9 : .. '0 - ' .I ?. LIJ 0. 0 U5 . ’ . _ V — O . i . 0 c1: ’ . ' '° . o O.| :- ' ‘7. 1' Z _. 4 _- "1 l l l l 4 5 6 7 (000/1‘ Figure 41. Plot of the 19F spin-lattice relaxation rate in CF2C1CF2C1 versus reciprocal temperature; the solid straight line represents the intramolecular dipolar relaxation rate and the dotted line represents the intermolecular dipolar relaxation rate in the solid phase. 205 The low temperature relaxation is evidently due to the intramolecular dipolar interaction modulated by the overall tumbling motion of the 286 molecules. From the geometry of the molecule, and assuming that the molecules are in the trans-form (which has been found to be more 292 stable than the gauche-form) we calculate the intramolecular dipolar relaxation rate according to Equation (17S) and obtain 8‘11 = 8.8X1091'r , (200) from which the reorientational correlation time Tr can be obtained; it is found to obey the equation 13exp(1.17 kcal mole'l/RT), (201) Tr = 4.6X10- which gives Tr=12 psec at the melting point. Due to the relatively slow reorientational motion, the spin- rotational relaxation is not important in the solid phase. Therefore, the curvature of the spin-lattice relaxation rates below the melting 'point must be due to the translational motion which shows an activation energy of ~9.5 kcal/mole. In the liquid state the molecular motion is more complicated due to the asymmetry of the molecules and the two possible conformers (trans and gauche) so the analysis was not attemped. 206 D. Molecular Motions in Some Other Plastic Crystals ~ In view of the similar plasticity found in CF3CC13 and in CC14, it is possible that the substitution of a trifluoromethyl group for a chlorine atom will not vary the physical characteristics of the perchlorocarbons very much. From Table 30 we noticed that the compounds obtained on substitution of the chlorines of C2C16 by CF3 groups (the last three compounds) still retain the high melting point and the small liquid range of C2C16. We also noticed that the physical appearance of these three compounds is very similar. They are all waxy, easy to sublime and having a camphor-like odor. We have studied the 19F NMR lineshapes and the relaxation behavior of these three compounds. Due to the more complex molecular structure, success in elucidating details of the molecular motions varies from compound to compound. 1. l,l,1-Trifluoropentachloropropane (TFPCP) This compound is formed by substituting one chlorine atom of C2Cl6 with a CF3 group. The measured 19F linewidths, the second moments and the spin-lattice relaxation rates are listed in Tables 43 and 44.. Figure 42 shows the plots of the 19F linewidths and the second moments versus reciprocal temperature. An obvious self-diffusion with a rate of the order of the linewidth occurs at ~-40°C, which is about 1400 below the melting point. M2d was found to be ~0.l45 G2 which can be 'compared with the theoretical value 0.105 G2 obtained from our measured density (1.712 g cm'3 at 25°C). From Figure 43,the spin-lattice relaxation rate in the solid phase shows a curvature, becoming a fairly straight line at low 207 9 F linewidths and second moments in CF3C12CCC13 Table 43. T 1000/T Av M2 (°C) (°x'1) (Hz) (62) 0.0 3.66 494 0.0068 ~8.0 3.77 693 0.0241 -16.0 3.89 1200 0.0408 -22.0 3.98 -- 0.0694 -27.0 4.06 1572 0.1150 -34.0 4.18 1954 0.1390 -37.0 4.24 3035 0.1430 -44.0 4.37 3056 0.148 29.0 3.31 132 0.0009 20.0 3.41 161 0.0020 13.0 3.50 210 0.0022 6.0 3.58 283 0.0130 0.0 3.66 459 - —7.0 3.76 770 - -15.0 3.88 1312 - -17.0 3.91 1655 - -23.0 4.00 2190 - -20.0 3.95 1950 - -32.0 4.15 2978 - -26.0 4.05 2820 - ~20.0 3.95 2160 - -15.0 3.88 1600 - -9.0 3.79 854 - Table 44 208 19F spin-lattice relaxation rates in CF3C12CCC13° T 1000/T R1 (°C) (°x'1) (sec'l) 23.0 3.38 0.1144i0.0037 16.0 3.46 0.122010.0022 8.0 3.56 0.1229i0.0042 0.0 3.66 0.1330t0.0034 -26.0 4.05 0.1556:0.0044 -20.0 3.95 O.1647i0.0051 -9.0 3.79 0.167010.0130 -15.0 3.88 0.1710:0.0070 32.0 3.28 0.1133i0.0007 39.0 3.20 0.1116i0.0030 40.0 3.19 0.1103:0.0008 49.0 3.10 0.1078i0.0018 56.0 3.04 0.110210.0014 64.0 2.97 0.1160:0.0023 65.0 2.96 0.1164i0.0020 86.5 2.78 0.142:0.002 95.0 2.72 0.162i0.003 102.0 2.67 0.161:0.004 97.0 2.70 0.17110.006 111.5 2.60 0.139i0.003 121.0 2.54 0.14010.004 146.0 2.39 0.151i0.007 173.0 2.24 0.17010.014 159.0 2.31 0.148:0.010 133.0 2.46 0.142:0.002 -42.0 4.33 0.28510.005 -43.0 4.35 0.310i0.009 -47.0 4.42 0.319:0.012 -37.0 4.24 0.277i0.027 -35.0 4.20 0.234i0.013 -30.0 4.12 0.197i0.007 -23.0 4.00 0.175:0.008 -18.0 3.92 0.159:0.006 -9.0 3.79 0.164t0.006 -l7.0 3.91 0.18110.006 -30.0 4.12 0.176:0.006 -24.0 4.02 0.226:0.007 209 0 Av . 1.5" 0 M2 " 0.3 A I L. 3 - 0.2 H (”A 5 2 3 0 x.) V i N <' 2 0.5- - 0.1 50 Figure 42. Plots of the 19F linewidths and the second moments in CF3C12CCC13 versus reciprocal temperature. 210 O. 5 '- " 1 005 .- l . ”I, . .1 L J l l I 2 3 4 IOOO/T Figure 43. Plot of the 19F_spin-lattice relaxation rate in CF3C12CCC13 versus reciprocal temperature; dashed line represents the intramolecular dipolar relaxation rate. 211 temperatures. The slope of this linear portion is ~2.74 kcal/mole, which can be reasonably assigned as the activation energy of the overall tumbling motion of the molecules. Assuming that the F-F internuclear distance in CFSCIZCCl3 is the same as that in CF3CC13, the intramolecular dipolar relaxation rate can be expressed by a? z 1.33X10101r (202) and the reorientational correlation time can be obtained as 1 T = S.SXl0- 4exp(2.74 kcal mole'l/RT), (203) 1‘ which gives Tr=2.2 psec at the melting point. The mean jump times estimated from the linewidth narrowing can be expressed by 13 z 4.8x10' exp(9.0 kcal mole-I/RT), (204) T6 which gives Td=8.le0-85ec at the melting point. Using the lattice diffusion model, this gives the self-diffusion coefficient to be -8 2 1 2 l.OXl0 cm sec' or l.lX10-8cm sec-E depending on whether the crystal structure is body—centered cubic or face-centered cubic. 2. 2,2, 3,3-Tfetrachlorohexafluorobutane (TCHFB) This compound is formed by substituting two chlorine atoms of sz6 by two CF3 groups. Weight and Roberts296 have studied the 19F 212 NMR spectra of this compound in CF2C12 solution. They found that at -112°C the equilibrium mixture contains 8% of the trans_and 92% of the gauche conformer. The activation energy for CF3 internal rotation was found to be 8 kcal/mole for the gaughg_isomer and 6 kcal/mole for the trans_isomer. They also concluded that the energy barrier restricting the rotation about the central carbon-carbon bond should be greater than 8 kcal/mole. We have measured the linewidths and the second moments of the 19F resonance lines in this compound, and also the spin-lattice relaxation rates. These are listed in Table 45. Plots of the linewidths and the second moments versus reciprocal temperature are shown in Figure 44. We did not observe a solid-solid transition but it may occur at a temperature lower than we attained (N-158°C). Both the linewidth and the second moment begin to decrease rapidly above -110°C,at which the self-diffusion reaches a speed of the order of the linewidth, ~5500 Hz. The second moment just before this decrease is ~0.25 6%,comparable to the theoretical value 0.19 G2 3 at 25°C). estimated from the density (~1.725 g cm' The spin-lattice relaxation rate is plotted in Figure 45. A maximum in the low temperature solid phase is unique. Since no transition point is detected, the low temperature relaxation is probably due to only one type of motion, namely, the overall molecular tumbling motion. To describe this motion, we first neglect the long range dipolar interaction of two CF groups since they are separated 3 by five bonds. (This assumption is valid for the trans isomer but might cause ~20% error for the gauche isomer.) Without using the fast- motion approximation, the intramolecular dipolar relaxation rates can 213 Table 45. 19F spin-lattice relaxation rates, linewidths and second moments in CF3CIZCCCIZCF3. T 1000/T R1 Av M2 (°C) (°x‘1) (sec‘l) (Hz) (a?) 29.0 3.31 0.11310.005 239 0.00085 9.0 3.55 0.834:0.004 581 0.0054 -59.0 4.67 0.47510.017 -- -- -74.0 5.02 0.75810.052 3015 0.115 -78.5 5.14 0.95410.088 3918 0.148 -91.S 5.49 1.62:0.11 4736 0.199 -100.0 5.78 3.32:0.10 5200 0.231 -107.5 6.04 4.80:0.28 5249 0.219 -115.0 6.33 7.26:0.39 5690 0.259 -122.0 6.62 10.820.6 5505 0.252 -128.0 6.90 12.010.8 5396 0.227 -l34.0 7.19 ll.ltO.S 5676 0.259 -141.0 7.58 8.4:0.2 -- -- -150.5 8.16 3.02:0.12 -- -- -158.0 8.70 0.65:0.10 -- -- -50.0 4.48 0.323:0.022 -- -- -32.5 4.16 0.190:0.022 -- -- -13.5 3.85 0.141:0.024 -- -- -4.0 3.58 0.104:0.001 -- -- 42.6 3.17 0.0964:0.0046 214 O : AV J LS)" O=-M2 0.3 LO" 5? m 3 <1 8 a < 0.5- l I I I 3 4 5 6 7 I000! T Figure 44. Plot of the 19F NMR linewidths and second moments in CF3C12CCC12CF3 versus reciprocal temperature. 215 l0— R|(5£c") - IOOO/T Figure 45. Plot of the 19F spin-lattice relaxation rate versus reciprocal temperature. 1n CF 3(212(3CC12CF3 216 be expressed by 4 _ T T 8’ =51 nzr 6( r + —-‘l—). (205) 1 5 1+m2T2 1+4m2T2 or 01' R: reaches a maximum when worr=0.6l6 and, from Figure 45, this occurs 0 at -130°c. With r=2.l6 A, Equation (205) gives (R{)max=10.6 sec'l, which is in very good agreement with the observed value. The difference ~1.6 see“1 is probably due to the additional contribution from the interaction between the two CF3 groups in the same molecule. At temperatures higher than -130°C, onr<<1 and Equation (205) can be simplified to 6 r 4 2 - R1 = 3 y M r Tr (206) and the reorientational correlation time T is found to be 1‘ 15exp(3.49 kcal mole-I/RT). (207) Tr = 7.9X10- The mean jump time can be estimated from the linewidth narrowing and we obtain -1 - Td = 1.96Xl0 2exp(8.3 kcal mole 1/RT), (208) -9 2 which gives Td=2.4SXI0-7sec,and DS=S.6XI0 cm sec"1 or 3.8x10'9cm2sec'l depending on whether the crystal has a body-centered or face-centered cubic structure. 217 3. 1,1,2,2-Tetrachloro-3,3,4,4-tetrafluorocyclobutane (TCTFCB) This compound can also be considered as a derivative of C2C16. But, instead of substituting chlorine atoms by CF3 groups, two chlorine atoms are substituted by a CFZ-CFZ group to form a cyclic compound. From the IR spectra of the vapor and of the crystalline phases, Harris 297 and Yang found that this compound has C2v symmetry, which means that the ring is planar. No crystal structure of this compound has been 29° found that an analogous compound, reported but Schapiro and Hoard l,2,3,4-tetrachlorotetrafluorocyclobutane, had a body-centered cubic crystal structure with lattice constant ~7.9510.1 X. We have measured the density of TCTFCB and find it is ~1.715 g cm‘3 at 25°C. If TCTFCB also has a body-centered cubic crystal structure, then the lattice constant will be ~8.02 R. The measured 19F linewidths, second moments, spin-lattice relaxation rates in the laboratory and in the rotating frames,and the spin-spin relaxation rates,are listed in Tables 46 to 48 and are plotted in Figures 46 to 48. The observed second moment, before its rapid decrease at tempera- tures above -25°C, is about 0.145 G2 which is in excellent agreement with the calculated value (0.142 62) from the lattice constant derived from the measured density. A possible phase transition at -67°C is observed based on the rapid increase of the second moment. The second moments obtained from the maxima of R1p are 0.14 G2, 0.123 G2 and 0.15 Gz,which agree very well with the above values. Figure 49 shows the claculated mean jump times for self-diffusion in the plastic phase of this compound from R10, R2 and the linewidth 218 Table 46. 19F spin~lattice relaxation rates, linewidths and second moments in 1,1,2,2-tetrachloro-3,3,4,4-tetraf1uoro- cyclobutane. T 1000/T R1 Av M2 (°C) (°K'1) (sec'l) (Hz) (62) -12.0 3.83 0.1098:0.0029 --- -- -18.0 3.92 0.1226:0.0030 928 0.0351 -24.0 4.02 0.1347:0.0039 1490 0.0595 -32.0 4.15 0.1640:0.0020 2396 0.0875 -39.0 4.27 0.180710.0060 3711 0.1325 -46.0 4.40 0.208010.0100 4273 0.1420 -53.0 4.54 0.2800:0.0180 4861 0.1530 -60.0 4.69 0.4060:0.0080 4882 0.1630 -ll.0 3.82 0.1140:0.0040 610 0.0222 -3.0 3.70 0.0975:0.0040 259 0.0032 7.0 3.57 -- 96 0.0005 6.0 3.58 0.0858:0.0025 -- -- 4.0 3.61 0.0953:0.0020 -- -- 13.0 3.50 0.0863:0.0012 74 0.0007 22.0 3.39 0.0837:0.0010 34 0.0001 31.0 3.29 0.080210.0015 20 0.0000 39.0 3.20 0.0846:0.0011 14 0.0000 55.0 3.05 0.085010.0020 12 0.0000 64.0 2.97 0.0835:0.0035 -- -- 71.0 2.91 0.096410.0019 -- -- 74.5 2.88 0.0887:0.0010 -- -- 79.0 2.84 0.0859:0.0016 -- -- 87.0 2.78 0.0878:0.0013 -- -- 98.0 2.70 0.0866:0.0021 -- -- 110.0 2.60 0.0888:0.0020 -- -- 87.0 2.78 0.0881:0.0009 -- -- 115.0 2.58 0.0908:0.0011 -- -- 142.0 2.41 0.1014:0.0024 -- -- 181.0 2.20 0.1220:0.0050 -- -- 218.0 2.04 0.1610:0.0060 -- -- 219 Table 47. 19F spin-lattice relaxation rates in the rotating frame in 1,1,2,2-tetrachloro-3,3,4,4-tetraf1uorocyclobutane. T (°C) 1000/T (°x'1) R1 (sec'l) H2 (kHz) -11.0 3.82 18.810.5 47.0 -21.5 3.98 9.0010.29 47.0 -27.0 4.06 8.86:0.25 47.0 -18.0 3.92 12.010.44 47.0 -14.0 3.86 16.710.30 47.0 -6.0 3.74 28.7510.99 47.0 -1.0 3.68 41.0:2.2 47.0 6.0 3.58 60.912.2 47.0 10.0 3.53 66.8:l.7 47.0 20.0 3.41 50.811.6 47.0 28.5 3.32 26.410.9 47.0 15.0 3.47 62.612.8 47.0 18.0 3.44 64.lil.3 47.0 24.5 3.36 31.9il.0 47.0 31.0 3.29 21.610.9 47.0 35.0 3.25 17.810.6 47.0 45.0 3.14 ll.3:0.3 47.0 55.0 3.05 6.6:0.1 47.0 22.0 3.39 37.7:0.7 47.0 20.0 3.41 54.810.5 25.0 2.0 3.64 11213 25.0 7.0 3.57 10412 25.0 13.0 3.50 8311 25.0 -4.0 3.72 10112 25.0 -l4.0 3.86 63.711.5 25.0 —21.0 3.97 36.810.9 25.0 -9.0 3.79 81.111.1 25.0 -29.0 4.10 21.3:0.8 25.0 -34.0 4.18 12.610.4 25.0 28.0 3.32 32.7:0.5 25.0 -34.0 4.18 40.6:6.2 12.5 -30.0 4.12 62.114.8 12.5 -23.0 4.00 10515 12.5 -17.0 3.91 15316 12.5 -12.0 3.83 19214 12.5 -8.5 3.78 25018 12.5 -4.0 3.72 26916 12.5 -1.0 3.68 260111 12.5 4.0 3.61 231110 12.5 9.0 3.55 170:4 12.5 12.0 3.51 14317 12.5 20.0 3.41 7S.9:1.l 12.5 33.0 3.27 31.510.7 12.5 26.0 3.34 53.210.9 12.5 220 Table 48. 19F spin-spin relaxation rates in l,l,2,2-tetrachloro- 3,3,4,4-tetrafluorocyclobutane. T 1000/T R2 (°C) (°x‘1) (sec‘l) -18.0 3.92 29141310 -11.0 3.82 18411150 -3.0 3.70 814140 0.0 3.66 410127 6.0 3.58 312122 8.0 3.56 269111 10.0 3.53 228111 15.0 3.47 136110 16.0 3.44 105111 22.0 3.39 58.814.0 31.0 3.29 49.312.7 32.0 3.28 40.712.7 41.0 3.18 20.610.9 50.0 3.10 12.410.4 58.0 3.02 7.4210.30 63.0 2.98 4.6610.20 64.0 2.97 4.5910.21 71.0 2.91 2.0310.08 87.0 2.78 0.086210.0029 115.0 2.58 0.10510.005 142.0 2.41 0.11310.004 181.0 2.20 0.13910.006 218.0 2.04 0.18510.012 221 oAV 0M2 0.15 O o ‘ ’ A 0 H | ~— “1 2 N83 0.1 0 L9 3 v g § v 4 N 2 .. e 0.5"“ 40.05 0 I. 'D I 3 a P Figure 46. Plot of the 19F NMR linewidths and second moments in O‘CIZCI’Z'CFZCCl2 versus reciprocal temperature. 222 O ' Iooot- C) m 1‘0 o 9 °\ I a - \ CI 0 \ o- ‘ loo— ,0 0. °\ . a ‘. \\ 4" ° °\ 0 .I \. \ _C‘ i .\ o G\ . -. \ 1‘11 " °\. Q U) ‘, \ E. . V ' I 0) °. ‘ 3.2 G CI P \.\ .I . a | 0_ ' 3. -6.3 6. I, h 9 e \ I2 G. 0_I I I I 3 4 Figure 47. 19F spin-lattice relaxation rates in the rotating frame at three different spin-locking fields (12, 6.3 and 3.2 G) and the spin-spin relaxation rate (0), in ({ClZCFZCFLC‘Cl2 . 223 0.02 r— ‘i ll '— l I ‘ I I h I— Figure 48. IOOO/T 19F spin-lattice relaxation rate (0) and spin-spin relaxation rate (A) in TCTFCB; solid circles are Rz-Rl; dashed line represents the intramolecular dipolar relaxation rate; dotted line represents the total spin-lattice relaxation rate after subtracting the intermolecular dipolar contribution. 224 I0 " .A i67~ IOOO/T Figure 49. The mean jump time for self-diffusion in the plastic phase of TCTFCB derived from R10 (0), R2 (0) and Av (A). 225 data. The best fit of these results can be expressed by -17 -1 Td = 4.BSXI0 exp(l3.9 kcal mole /RT), (209) which gives rd21.67x10’°sec,and therefore D524.SXI0-°cm25ec'l,at the melting point. To analyze the spin-lattice relaxation data, we assume that the molecules are planar. Using the standard bondlengths and bond angleszg? the intramolecular dipolar relaxation rate can be expressed by d 9 ' R1 = 8.74X10 Tr. (210) The reorientational correlation time Tr can then be obtained from the low temperature R1 values. We obtain - -1 Tr = 2.18X10 14eXp(3.18 kcal mole /RT), (211) which gives Tr=2.05 psec at the melting point. The intermolecular dipolar relaxation rate in the plastic solid phase can be estimated from the mean jump time for self-diffusion. This can be expressed by the relation 5 -9 R1 - 1.64x10 /Td. (212) In Figure 48, we subtract this contribution from the total spin- lattice relaxation rate. The remaining relaxation rate is represented I by the dotted line. An important contribution from the spin-rotational 226 interaction is very obvious even below the melting point. We have observed that in the liquid phase of TCTFCB the spin-spin relaxation rate also has an additional contribution from the scalar coupling due to the fast-relaxing chlorine nuclei. Although this con— SC1) tribution is small, we can still estimate the coupling constant J(F3 from it. Using the same approach used in the previous sections, we obtain J(FSSCI) to be ~1.710.2 Hz. E. Discussion 1. Rotational Motion in Plastic Crystals One of the characteristic properties of plastic crystals is the rotational freedom of molecules in the plastic solid phase. In Table 49, we summarize the motional parameters of the plastic crystals we studied in this work and some other compounds studied elsewhere243’246. From the last column of this table, it can be easily seen that the reorientational correlation times of molecules in the plastic crystals are all of the order of picoseconds,as normally found in the liquid phase. This implies that the molecules in the plastic phase have already obtained complete freedom of reorientation. In non-plastic crystals, e.g., hexamethyltetramine (HMT), Tr just below the melting point is much longer. In CFZCICFZCl, which we expected to have little plasticity, It just below its melting point is also longer than the order of picoseconds. 227 Table 49. Motional parameters of some plastic crystals. Compound AEr AEt AEi.r. 108Ds Tr (kcal mole‘l) (cmzsec'l) (psec) 053cm3 1.8c 11.7 1.26b 1.61 2.53c 12.8C 1.5c cpzc1c013 2.56 11.6 4.1 1.0 2.25 12. 8g CFCIZCFCIZ 3.16 11.4 9.0f 1.5 1.21 CF c01 cc1 2.74 39.0 5.7d 0.16 2.2 3 2 3 e CF3C12CCC12CF3 3.49 28.3 8 0.36 1.1 - 5 -- j _- j -- c C6F12 14.1 0.2 CwZCCIZCCIZpFZ 3.18 13.9 -- 4.8 2.05 k k k k Adamantanek 3.08 36. .72 -- 0.1 1.65 33.0 CFZClCFZCIi 1.17 9.5 —- -- 12.0 k Hexamethyl- 19.3k -- -- 10"s 7641k tetraminei aInternal rotation of CFn group in the molecule. bValues at the melting point. cReference 131. dIn CF C12 solution, F. J. Weigert and W. Mahler, J. Am. Chem. Soc. 94, 5314 (1972). eFor the gauche isomer, F. J. Weigert and J. D. Roberts, J. Am. Chem. Soc. 20, 3577 (1968). fIn CFC13 solution, Reference 287. 8N. w. Lufts, J. Phys. Chem. 52, 92 (1955). hReference 289. iNon-plastic crystal. jReference 243. 1‘Reference 246. 2E. Hampton, N. C. Lockhart and J. N. Sherwood, Chem. Phys. Lett. 21, 191 (1973). 228 2. Self-Diffusion in Plastic Crystals Comparison of self-diffusion studies from NMR spectroscopy, and from other methods such as radiotracer and plastic flow, has been the aim of a series of papers published by Sherwood and his coworker5238-242 and by Strange and his coworker5233’237. They found that the activation energy of the self-diffusion obtained from NMR relaxation studies (Afigmr) is not always equal to that from radiotracer or plastic de- formation work (A82). Since both the radiotracer and plastic deformation techniques have a longer time scale, the measured diffusion constant is that for the true macroscopic self-diffusion. 0n the contrary, the NMR eXperiment has a much smaller time scale and therefore should reflect both the microscopic and the macroscopic diffusive motions. In those solids with high entrapy of fusion (about 2R), e.g., hexamethyl- ethane or perfluorocyclohexane, Afigmr=AEg=2Ls, where L5 is the latent heat of sublimation. This result is consistant with the vacancy diffu- sion mechanism in which both the formation and the migration of a vacancy involve an entropy change of magnitude Ls. 0n the other hand” Afigmr was found to besmaller than A52 for those plastic crystals with a low entropy of fusion (about R). In these crystals, the defect which caused diffusion might involve many molecules (12 to 20) and form a small disordered region in the lattice or a vacancy into which the surrounding molecules have relaxed. Since insufficient thermodynamic data are available for the compounds studied here, we are not able to make the same comparison. However, we find that the rate of the self-diffusive motion in the plastic solid phase near the melting point is about the same order of 229 magnitude for all the plastic crystals we studied and also for some 300 243’246. Recently, Boden gt_gl: have discussed the studied by others failure of the Torrey model for describing self-diffusion in some plastic crystals with a low entropy of fusion. They found that for these crystals the isotropic diffusion model, which does not restrict the mean jump distance to be equal to the lattice constant, gives a better result. The main difference in using the isotropic diffusion model is the greater Td (by about a factor of two) at temperatures near the melting point than that predicted from Torrey's lattice-diffusion model. For the compounds we studied, the Torrey model seems to be appropriate and so is used. 3. Test of the Theory of Fusion Smith232 has tried to relate the parameter T, defined as the ratio of the rotational to the translational activation energies, to the parameter v in the Peple-Karasz theory of fusion41 as modified by Amzel and Becka42 . In Table 50, we summarize the comparison of the parameters r and v and the physical constants of the plastic crystals we studied. From this table, it can be easily seen that P and v are very well correlated. Table 50. Comparison of the parameters F and 230 v in some plastic crystals. Compound T$a TEa’d Pb c CFSCCl3 0.65 0.33 0.14 .13 CF2C1CC13 0.62 0.37 0.21 .20 CFC12CFC12 0.59 (0.41) 0.28 .30 CFSClZCCC13 0.64 (0.34) $0.30 .16 CFSCIZCCCIZCFS 0.64 (0.34) $0.42 .16 CFZCC12CCIZEF2 0.64 0.34 0.18 .16 aT* = 0.72 T/Tb. m: melting; t: b r = AEr/AES. transition; b: boiling. CPredicted from T5 value according to the Pople-Karasz theory. dValues in parenthesis are predicted from Reference 232. SUMMARY A general program has been written to perform a large variety of pulsed NMR studies automatically. Several interfaces were constructed to permit varying the pulsewidths, to supply a constant pulsed current to the field-gradient coil and to alternate the radiofrequency pulses. Complete 14N and 19 F relaxation studies of CF3CN were accomplish- ed and the motional anisotropy of the molecules in the liquid state was determined. The J-diffusion model was found to be appropriate to describe the reorientational motion of the CF3CN molecules. Several physical constants of the molecules were determined, e.g., the chemical 13 shift anisotrOpy and the spin-rotation constants for both the C and the 19F nucleus. 13C relaxation studies have been carried out for CDBr3, CFBr3 and CHZBrZ and the scalar coupling constants between 13C and 798r determined from the dominant scalar relaxation rates in these highly brominated compounds. In CHZBr2 the coupling constant J(C7gBr) was found to be 58.7 Hz, a reasonable value compared with the reported value for J(CSSCI). But,in both CDBr and CFBr3,J(C7gBr) was found 3 to be about twice as big as expected and this discrepancy has been discussed. The coupling constant J(F798r) in CFBr3 was also deter- mined and found to be 30.5 Hz from the 19F relaxation studies. The test of the extended diffusion model in this molecule has been dis- cussed. An attempt at resolving the anisotropic motions of planar CFZCCl2 molecules in the liquid state has been made with a certain 231 232 degree of success.It was found that the reorientational motion in the liquid and in the solid states was via different mechanisms. The separation of various 19F relaxation mechanisms in CF3I was performed but difficulties were found in extracting the intramolecular dipolar relaxation rates and therefore analysis of the reorientational motion was fruitless. However, it was found that the spin~rotationa1 relaxation mechanism was via the spinning motion of the molecules about the C3 axis. A number of plastic crystals were studied by analysis of the 19F NMR lineshapes and by measurement of the relaxation rates of the 19F nuclei. Measurement of the 19F spin-lattice relaxation rate in the rotating frame was found to be very informative about the self- diffusive motion in the plastic phase of these crystals. The mean jump time between adjacent diffusion steps was determined to several orders of magnitude. In all the plastic crystals studied the Torrey lattice-diffusion model was found to be appropriate to describe the self-diffusion of the molecules in the plastic solid phase. The reorientational motion in these plastic crystals was found to be as fast as in the liquid phases. From the 19F spin-spin relaxation rates, the scalar-coupled relaxation rates in someof'these fluorochlorocarbon derivatives were extracted and the coupling constants J (F35Cl) determined. Finally, a test of the theory of fusion was discussed in terms of the activation parameters derived from the NMR studies. ’ RBFBRENC ES 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. REFERENCES For a short review of work on master equations up to 1964, see R. Zwanzig, Physica, 39, 1109 (1964). For a review of the extensions of the master equation, see N. G. van Kampen, Adv. Chem. Phys. 35, 245 (1976). R. G. Gordon, Adv. Mag. Res. 3, l (1969). R. Zwanzig, Ann. Rev. Phys. Chem. 33, 67 (1965). L. van Hove, Phys. Rev. 23, 249 (1954). For example, see H. Conroy, J. Chem. Phys. 33, 1331 (1964). B. J. Alder and T. E. Wainwright, J. Chem. Phys. 33, 459 (1959); 33, 1439 (1960). A Rahman, Phys. Rev. A136, 405 (1964). L. Verlet, Phys. Rev. 159, 98 (1967). B. J. Berne and G. D. Harp, Adv. Chem. Phys. 31, 63 (1970). H. C. Longuet-Higgins and J. A. Pople, J. Chem. Phys. 33, 884 (1956). 0. E. O'Reilly, M01. Phys. 39, 1189 (1975). H. T. Davis, S. A. Rice and J. V. Sengers, J. Chem. Phys. 33, 2210 (1961). R. Brown and H. T. Davis, J. Phys. Chem. 33, 1970 (1971). H. C. Anderson, D. Chandler and J. D. Weeks, Adv. Chem. Phys. 31, 105 (1976). P. Debye, "Polar Molecules", Dover, New York (1929). C. Brot and I. Darmon, Mel. Phys. 33, 785 (1971). R. I. Cukier and K. Lakatos-Lindenberg, J. Chem. Phys. 31, 3427 (1972). K. Lakatos-Lindenberg and R. I. Cukier, J. Chem. Phys. 33, 3271 (1975). B. J. Berne, J. Chem. Phys. 33, 1154 (1975). 233 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 234 N. A. Steele, Adv. Chem. Phys. 33, l (1976). R. G. Gordon, J. Chem. Phys. 33, 1830 (1966). R. E. D. McClung, J. Chem. Phys. 33, 3842 (1969); 33, 3248 (1971); 33, 3459 (1971); 33, 5478 (1972). M. Fixman and K. Rider, J. Chem. Phys. 33, 2425 (1969). A. G. St. Pierre and W. A. Steele, J. Chem. Phys. 33, 4638 (1972). F. Bliot, C. Abbar and E. Constant, Mol. Phys. 33, 241 (1972). F. Bliot and E. Constant, Chem. Phys. Lett. 33, 253 (1973). F. Bliot and E. Constant, Chem. Phys. Lett. 33, 618 (1974) R. E. D. McClung, Advances in Molecular Relaxation and Interaction Processes, 39, 83 (1977). W. A. Steele, J. Chem. Phys. 33, 2404 (1963); 2411 (1963). P. W. Atkins, MOI. Phys. 33, 321 (1969). F. A. Lindemann, Physik Z. 33, 609 (1910). R. H. Brout, "Phase Transitions", W. A. Benjamin Inc.,'New York (1965). L. L. Lee, J. Chem. Phys. 33, 4436 (1975). B. Jancovici, Physica, 33, 1017 (1965). .J. A. Barker and D. Henderson, Mol. Phys. 33, 587 (1968). J. E. Lennard-Jones and A. F. Devonshire, Proc. Roy. Soc. A169, 317 (1939); A170, 464 (1939). L. K. Runnels, Phys. Rev Lett. 33, 581 (1965); L. K. Runnels and L. L. Combs, J. Chem. Phys. 33, 2482 (1966); L. K. Runnels, L. L. Combs and J. P. Salvant, J. Chem. Phys. 33, 4015 (1967). A. Bellemans and R. K. Nigam, Phys. Rev. Lett. 33, 1038 (1966); J. Chem. Phys. 33, 2922 (1967). A. Bellmans and A. Orban, Phys. Rev. 33, 908 (1966); Chem. Phys. Lett. 3, 205 (1967); J. Orban, J. van Craen and A. Bellmans, J. Chem. Phys. 33, 1778 (1968). J. A. Pople and F. B. Karasz, J. Phys. Chem. Solids, 33, 28 (1961); 33, 294 (1961). L. M. Amzel and L. N. Becka, J. Phys. Chem. Solids, 39, 521 (1969). 43. 44. 45. 46. 47. 48. 49. SO. 51. 52. S3. 54 SS. 56. S7. 58. $9. 60. 61. 62. 63. 235 C. J. Thomson, "Mathematical Statistical Mechanics", P. W. Kasteleyn, New York (1973). E. M. Purcell, H. C. Torrey and R. V. Pound, Phys. Rev. Q, 37 (1946). F. Bloch, W. W. Hansen and M. Packard, Phys. Rev. 99, 127 (1946). See, for example, E. Merzbacher, "Quantum Mechanics", 2nd Ed., John Wiley and Sons Inc., New York (1970). F. Bloch, Phys. Rev. 39, 460 (1946). R. K. Wangsness and F. Bloch, Phys. Rev. 39, 728 (1953). F. Bloch, Phys. Rev. 393, 104 (1954); 393, 1206 (1957). A. G. Redfield, Adv. Mag. Res. 3, 1 (1965). A. Carrington and A. McLachlan, "Introduction to Magnetic Resonance", Harper and Row, New York (1967). N. Bloembergen, E. M. Purcell and R. V. Pound, Phys. Rev. 33, 679 (1948). H. C. Torrey, Phys. Rev. 93, 962 (1953) A. Abragam, "Principle of Nuclear Magnetism", Oxford UP, New York (1961). H. G. Hertz, Progress in NMR Spectroscopy, 3, 159 (1967). W. T. Huntress, J. Chem. Phys. 33, 3524 (1968). R. R. Sharp, J. Chem. Phys. 99, 1149 (1974). H S. Gutowsky, I. J. Lawrenson and K. Shimomura, Phys. Rev. Lett. 3, 349 (1961); H. S. Gutowsky and D. E. Woessner, Phys. Rev. 104, 843 (1956); R. J. C. Brown, H. S. Gutowsky and K. Shimomura, J. Chem. Phys. 39, 76 (1963). C. S. Johnson, Jr., J. S. Waugh and J. N. Pinkerton, J. Chem. Phys. 33, 1128 (1961). C. S. Johnson, Jr. and J. S. Waugh, J. Chem. Phys. 33, 2020 (1961). P. S. Hubbard, Phys. Rev. 131, 1155 (1963). C. H. Wang, D. M. Grant and J. R. Lyerla, Jr., J. Chem. Phys. 33, 4674 (1971). C. H. Wang, J. Magn. Resonance 9, 75 (1973). 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 236 C. H. Wang, Mol. Phys. 39, 801 (1974). L. D. Favro, Phys. Rev. 339, 53 (1960). D. E. Woessner, J. Chem. Phys. 33, 647 (1962). H. Shimizu, J. Chem. Phys. 33, 765 (1962); 39, 754 (1964). W. T. Huntress, Jr., Adv. Mag. Res. 3, 1 (1970). J. W. Cooper, "The Computer in FTNMR" in "Topics in Carbon-13 NMR" Vol. 2, G. Levy, Ed., John Wiley and Sons, New York (1975); J. W. C00per, Computers and Chemistry, 3, 55 (1975); J. Magn. Resonance 22, 345 (1976); J. W. Cooper, "The Computer in the Laboratory", Wiley, New York (1976). See, for example, S. Goldman, "Information Theory", Dover, New York (1968). R. R. Ernst and W. A. Anderson, Rev. Sci. Instrum. 33, 93 (1966). R. M. Bracewell, "The Fourier Transform and its Application", McGraw Hill, New York (1965). J. D. Ellett, M. G. Gibby, U. Haeberlin, L. M. Huber, M. Mehring, A. Pines and J. S. Waugh, Adv. Mag. Rev. 3, 117 (1971). A. G. Redfield and R. K. Gupta, Adv. Mag. Res. 3, 82 (1971). E. 0. Stejskal and J. Schaefer, J. Magn. Resonance 33, 249 (1974); 33, 160 (1974). A. G. Redfield and S. D. Kunz, J. Magn. Resonance 39, 250 (1975). D. I. Hoult, J. Magn. Resonance, 33, 337 (1976). R. T. Pajer and I. M. Armitage, J. Magn. Resonance 33, 365 (1976). S. 1. Parks and R. B. Johannesen, J. Magn. Resonance 33, 265 (1976). W. K. Rhim, D. P. Burum and R. W. Vaughan, Rev. Sci. Instrum. 33, 720 (1976). S. Schaublin, A Hohener and R. R. Ernst, J. Magn. Resonance 33, 196 (1974). R. L. Vold, J. S. Waugh, M. P. Klein and D. E. Phelps, J. Chem. Phys. 39, 3831 (1968). J. L. Markley, W. J. Horsley and M. P. Klein, J. Chem. Phys. '33, 3604 (1971). 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 237 R. Freeman and H. D. W. Hill, J. Chem. Phys. 33, 3367 (1971); R. Freeman, H. D. W. Hill and R. Kaptein, J. Magn. Resonance 3, 82 (1972). D. Canet, G. C. Levy and I. R. Peat, J. Magn. Resonance 39, 199 (1975). R.K8Pteifl. K. Dijkstra and C. E. Tarr, J. Magn. Resonance _23, 295 (1976) . R. K. Gupta, J. Magn. Resonance 33, 231 (1977). Y. N. Luzikov, N. M. Sergeyev and M. G. Levkovitch, J. Magn. Resonance 33, 359 (1976). J. Kronenbitter and A. Schwenk, J. Magn. Resonance 33, 147 (1977). D. E. Demco, V. Simplaceanuand I. Ursu, J. Magn. Resonance 33, 166 (1974). ' R. J. Kurland and R. G. Parrish, J. Magn. Resonance 33, 295 (1975). H. T. Edzes, J. Magn. Resonance 33, 301 (1975). T. Cosgrove, J. Magn. Resonance 33, 469 (1976). G. C. Levy and I. R. Peat, J. Magn. Resonance 39, 500 (1975). R. Freeman and H. D. W. Hill, J. Chem. Phys. 33, 985 (1971). R. L. Vold and S. 0. Chan, J. Magn. Resonance 3, 208 (1971). R. L. Vold, J. Chem. Phys. 39, 3210 (1972). H. Haeberlen, H. W. Spiess and D. Schweitzer, J. Magn. Resonance 9, 39 (1972). R. L. Vold and S. 0. Chan, J. Chem. Phys. 39, 28 (1972). A. C. McLaughlin, G. G. McDonald and J. S. Leigh, Jr., J. Magn. Resonance 33, 107 (1973). R. L. Vold, R. R. Vold and H. E. Simon, J. Magn. Resonance 33, 283 (1973). R. L. Vold and R. R. Vold, J. Chem. Phys. 93, 2525 (1974). R. L. Vold, R. R. Vold and H. E. Simon, J. Magn. Resonance 33, 226 (1974). L. G. Werbelow and D. M. Grant, J. Magn. Resonance 39, 554 (1975). 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 238 R. R. Vold and R. L. Vold, J. Chem. Phys. 93, 320 (1976). A. Kumar and R. R. Ernst, J. Magn. Resonance 33, 425 (1976). R. Freeman and H. D. W. Hill, "Dynamic NMR Spectroscopy", Ed. by L. M. Jackman and F. A. Cotton, Academic Press, New York (1975). 1. Solomon, Phys. Rev. 99, 559 (1955). J. H. Noggle, J. Chem. Phys. 33, 3304 (1965). P. S. Hubbard, J. Chem. Phys. 33, 3546 (1965); J. H. Chaffin and P. S. Hubbard, J. Chem. Phys. 39, 1511 (1967); P. S. Hubbard, J. Chem. Phys. 33, 1647 (1969). L. G. Werbelow and A. G. Marshall, J. Magn Resonance 33, 299 (1973); Mol. Phys. 39, 113 (1974). I. D. Campbell and R. Freeman, J. Magn. Resonance 33, 143 (1973). L. G. Werbelow and D. M. Grant, J. Chem. Phys. 93, 544 (1975). C. L. Mayne, D. W. Alderman and D. M. Grant, J. Chem. Phys. 93, 2514 (1975). P. E. Fagerness, D. W. Alderman, K. F. Kuhlmann, C. L. Mayne and R. B. Bray, J. Chem. Phys. 93, 2524 (1975). C. H. Wang and D. M. Grant, J. Chem. Phys. 93, 1522 (1976). L. G. Werbelow and D. M. Grant, J. Chem. Phys. 93, 4742 (1975). C. L. Mayne, D. M. Grant and D. W. Alderman, J. Chem. Phys. 93, 1684 (1976). J. M. Courtieu, C. L. Mayne and D. M. Grant, J. Chem. Phys. 99, 2669 (1977). H. Y. Carr and E. M. Purcell, Phys. Rev. 93, 630 (1954). S. Meiboom and D. Gill, Rev. Sci. Instrum. 39, 688 (1958). R. L. Streever and H. Y. Carr, Phys. Rev. 121, 20 (1961). Dinesh, M. T. Rogers and G. D. Vickers, Rev. Sci. Instrum. 33, 555 (1972). Dinesh and M. T. Rogers, J. Chem. Phys. 56, 542 (1972). D. D. Traficante, J. A..Simms and M. Mulcay, J. Magn, Resonance 33, 484 (1974). 126. 127. 128. 129. 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. 147. 239 T. C. Farrar and E. D. Becker, "Pulse and Fourier Transform NMR", Academic Press, New York and London (1971). W. G. Clark, Rev. Sci. Instrum. 33, 316 (1964). I. J. Lowe and C. E. Tarr, J. Phys. E, Series 2, 3, 320 (1968). R. A. McKay and D. E. Woessner, J. Sci. Instrum. 33, 838 (1966). D. I. Hoult and R. E. Richards, J. Magn. Resonance 33, 71 (1976). D. A. Wright, Ph.D. Thesis, Michigan State University, East Lansing, Michigan (1974). G. G. McDonald and J. S. Leigh, J. Magn. Resonance 9, 358 (1973). S. G. Huang and M. T. Rogers, Chem. Instrum. 9, 17 (1977). E. 0. Stejskal and J. E. Tanner, J. Chem. Phys. 33,288 (1965). J. E. Tanner, Rev. Sci. Instrum. 39, 1086 (1965). J. E. Tanner and E. O. Stejskal, J. Chem. Phys. 39, 1768 (1968). J. E. Tanner, J. Chem. Phys. 33, 2523 (1970). K. J. Packer, C. Rees and D. J. Tomlinson, Mol. Phys. 39, 421 (1970). R. Blinc, J. Pirs and I. Zupancic, Phys. Rev. Lett. 39, 546 (1973). A. Pines and T. W. Shattuck, Chem. Phys. Lett. 33, 614 (1973). T. L. James and G. G. McDonald, J. Magn. Resonance 33, 58 (1973). J. S. Murday, J. Magn. Resonance 39, 111 (1973). B. W. Bangerter, J. Magn. Resonance 33, 87 (1973). G. Odberg and L. Odberg, J. Magn. Resonance 39, 342 (1974). D. S. Webster and K. H. Marsden, Rev. Sci. Instrum. 33, 1232 (1974). R. A. Assink, J. Magn. Resonance 3;, 165 (1976). J. Andrasko, J. Magn. Resonance 33, 479 (1976). 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163. 164. 165. 166. 167. 240 K. F. Kuhlmann and D. M. Grant, J. Am. Chem. Soc. 99, 7355 (1968); K. F. Kuhlmann, D. M. Grant and R. K. Harris, J. Chem. Phys. 33, 3439 (1970). D. Doddrell, V. Glushko and A. Allerhand, J. Chem. Phys. 39, 3683 (1972). .J. B. Stothers, "Carbon-l3 NMR Spectroscopy", Academic Press, New York (1972). G. C. Levy and G. L. Nelson, "Carbon-13 NMR for Organic Chemists", Interscience (1972). R. Freeman, H. D. W. Hill and R. Kaptein, J. Magn. Resonance 3, 327 (1972). J. R. Lyerla, Jr., D. M. Grant and R. D. Bertrand, J. Phys. Chem. 33, 3967 (1971). W. Spiess, D. Schweitzer, U. Haeberlen and K. H. Hausser, Magn. Resonance 3, 101 (1971). L4: R. J. Schoenberger, R. B. Creel, B. K. Lunds and R. G. Barnes, J. Magn. Resonance 33, 20 (1973). A ( . Kumar and C. S. Johnson, Jr., J. Magn. Resonance 3, SS 1972). R. A. Assink, J. Magn. Resonance 33, 88 (1973). K. van Putte, J. Magn. Resonance 3, 174 (1970). T. E. Bull, Rev. Sci. Instrum. 33, 232 (1974). D. E. Demco, V. Simplaceanu and I Ursu, J. Magn Resonance 33, 310 (1974). J. D. Cutnell, H. E. Bleich and J. A. Glasel, J. Magn. Resonance 33, 43 (1976). T. T. BopP. J. Chem. Phys. 33, 3621 (1967). D. E. Woessner, B. S. Snowden, Jr. and E. T. Strom, Mol. Phys. 33, 269 (1968). K. T. Gillen, M. Schwartz and J. H. Noggle, Mbl. Phys. 39, 899' (1971). K. T. Gillen and J. H. Noggle, J. Chem. Phys. 33, 801 (1970). J. Sheridan and W Gordy, J Chem. Phys. 39, S91 (1952). W. G. Mbniz and H. S. Gutowsky, J. Chem. Phys. 38, 1155 (1963). 168. 169. 170. 171. 172. 173. 174. 175. 176. 177. 178. 179. 180. 181. 182. 183. 184. 185. 186. 187. 188. 189. 241 K. T. Gillen and J. H. Noggle, J. Chem. Phys. 33, 4905 (1970). J. M. Lehn and J. P. Kintzinger, "Nitrogen NMR", M. Witanowski and G. A. Webb, Ed., Plenum Press, New York (1973). K. “r. Gillen, J. Chem. Phys. 96, 1573 (1972). K. T. Gillen, D. C. Douglass, M. S. Malmberg and A. A. Maryott, J. Chem. Phys. 33, 5170 (1972). W. T. Huntress, Jr., J. Phys. Chem. 33, 103 (1969). T. E. Bull, J. Chem. Phys. 93, 222 (1975). F. J. Bartoli and T. A. Litovitz, J. Chem. Phys. 39, 413 (1972). J. E. Griffiths, J. Chem. Phys. 39, 751 (1973). G. D. Paterson and J. E. Griffiths, J. Chem. Phys. 93, 2406 (1975). K. Mansingh and A Mansingh, J. Chem. Phys. 33, 827 (1964). A. A. Maryott, T. C. Farrar and M. S. Malmberg, J. Chem. Phys. 33, 64 (1971). J. H. Rugheimer and P. S. Hubbard, J. Chem. Phys. 39, 552 (1963). W. R. Hackleman and P. S. Hubbard, J. Chem. Phys. 39, 2688 (1963). P. Rigny and J. Virlet, J. Chem. Phys. 33, 4645 (1967). J. C. Hindman, J. G. Malm, A. Suirmickas and H. D. Frame, J. Magn. Resonance 9, 428 (1973). J. H. Chaffin III and P. S. Hubbard, J. Chem Phys. 46, 1511 (1967). J. W. Harrell, Jr., J. Magn. Resonance 33, 335 (1976). W. H. Flygare, J. Chem. Phys. 33, 793 (1964). C. Deverall, Mol. Phys. 39, 319 (1964) N. F. Ramsey, Phys. Rev. 39, 699 (1950). D. K. Hinderman and C. D. Cornwell, J. Chem. Phys. 39, 4153 (1968). R. A. Bonham and T. G. Strand, J. Chem. Phys. 39, 3447 (1964). 190. 191. 192. 193. 194. 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. 205. 206. 207. 208. 209. 242 D. Wallach and W. T. Huntress, Jr., J. Chem. Phys. 39, 1219 (1969). R. E. D. MCClung, Advances in Melecular Relaxation and Interaction Processes, 39, 83 (1977),and references therein. A. J. Montana, B. R. Appleman and B. P. Dailey, J. Chem. Phys. 99, 1850 (1977) J. A. Faniran and H. F. Shaurvell, Spectrochim. Acta, A27, 1945 (1971). W. M. Litchman and D. M. Grant, J. Am. Chem. Soc. 99, 1400 (1968). J. A. Paple, W. G. Schneider and H..J. Bernstein, "High Resolution Nuclear Magnetic Resonance", McGraw Hill, New York (1959). N. Muller and D. T. Carr, J. Phys. Chem. 93, 112 (1963). R. A. DeMarco, W. B. Fox, W. B. Mbniz and S. A. Sojka, J. Magn. Resonance. 39, 522 (1975). D. A. Wright and M. T. Rogers, J. Chem. Phys. 63, 909 (1975), R. Shoup and T. C. Farrar, J. Magn. Resonance 3, 48 (1972). R. Q. Williams, J. T. Cox and W. Gordy, J. Chem. Phys. 39, 1524 (1952). H. Zeldes and R. Livingston, J. Chem. Phys. 33, 1418 (1953). A. Briguet, J. C. Duplan and J. Delmau, Mel. Phys. 39, 837 (1975). T. C. Farrar, S. J. Druck, R. R. Shoup and E. D. Becker, J. Am. Chem. Soc. 93, 699 (1972). R. C. Miller and C. P. Smyth, J. Chem. Phys. 33, 814 (1956). H. Moissan and 0. Ruff, "Fluorine Chemistry", J. H. Simona, Ed., Vol. 1, Academic Press, New York (1950). A. Di Giacomo and C. P. Smyth, J. Am. Chem. Soc. 33, 774 (1955). H. S. Sandhu, J. Magn. Resonance 39, 7 (1977). P. M. Brown, N. R. Krishna and S. L. Gordon, J. Magn. Resonance 39, 540 (1975). C. H. Wang, D. R. Jones and D. H. Christensen, J. Chem. Phys. 93, 2820 (1976). 243 210. C. Hu and R. Zwanzig, J. Chem. Phys. 99, 4345 (1974). 211. A. Gierer and K. Wirtz, Z. Naturforsch. 39, 532 (1953). 212. R. Freeman, R. R. Ernst and W. A. Anderson, J. Chem. Phys. 39. 1125 (1967). 213. D. Wallach and W. T. Huntress, Jr., J. Chem. Phys. 39, 1219 (1969). 214. P. D'Antonic and C. George, J. Chem. Phys. 93, 2884 (1976). 215. G. V. D. Tiers and P. C. Lauterbur, J. Chem. Phys. 39, 1110 (1962). 216. R. Kaiser and A. Saika, Mol. Phys. 33, 221 (1968). 217. J. W. Emsley and L. H. Sutcliffe, Progress in NMR Spectroscopy, 39, 314 (1976). 218. J. A. Howe, J. Chem. Phys. 33, 1247 (1961). 219. R. Livingston, J. Phys. Chem. 33, 496 (1953). 220. F. Sterzer, J. Chem. Phys. 33, 762 (1955). 221. A. L. Andreassen and S. H. Bauer, J. Chem. Phys. 39, 3802 (1972). 222. F. Simon and C. van SimOn, Z. Phys. 33, 168 (1924). 223. L. Pauling, Phys. Rev. 39, 430 (1930). 224. J. Timmermans, J. Chin. Phys. 33, 331 (1938). 225. For a historical review, see (a) J. Phys. Chem. Solids, 18, No. 1 (1961); (b) J. G. Aston, "Physical Chemistry of the Organic Solid State", p.543, Labes, Weissberger and Fox, Ed., Wiley (1963). 226. L. A. K. Staveley, J. Phys. Chem. Solids, 39, 46 (1961). 227 C. P. Smyth, J. Phys. Chem. Solids 39, 80 (1961). 228. W. J. Dunning, J. Phys. Chem. Solids 39, 21 (1961). 229. A. Michils, Bull. Soc. Chim. Belg. _5_7, 575 (1948). 230. N. C. Lockhart and J. N. Sherwood, Faraday Symp. Chem. Soc. 9_57 (1972). 231. E. R. Andrew and P. S. Allen, J. Chem. Phys. 93, 85 (1966). 232. 233. 234. 235. 236. 237. 238. 239. 240. 241. 242. 243. 244. 245. 246. 247. 248. 249. 250. 251. 244 c. N. Smith, J. Chem. Phys. 39, 3081 (1962); 53, 4229 (1965); 4_3, 4325 (1965); fl, 3595 (1969). J. M. Chezeau, J. Dufourcq and J. H. Strange, M01. Phys. 39, 305 (1971). R. L. Jackson and J. H. Strange, M01. Phys. 33, 313 (1971). R. Folland, R. L. Jackson, J. H. Strange and A. V. Chadwick, J. Phys. Chem. Solids 33, 1713 (1973). R ( . Folland, S. M. Ross and J. H. Strange, Mel. Phys. 39, 27 1973). S. M. Ross and J. H. Strange, Mol. Cryst. Liq. Cryst. 39, 321 (1976). H. A. Resing, N. T. Corke and J. N. Sherwood, Phys Rev. Lett. 39, 1227 (1968). H. M. Hawthorne and J. N. Sherwood, Trans. Faraday Soc. 99, 1783 (1970); 1792 (1970); 1799 (1970). H. Blum, N. C. Lockhart and J. N. Sherwood, Mol. Phys. 39, 577 (1971). H. Blum and J. N. Sherwood, Mol. Cryst. Liq. Cryst. 39, 381 (1970). P. Bladon, N. C. Lockhart and J. N. Sherwood, Mol. Cryst. Liq. Cryst. 39, 315 (1973). N. Boden, J. Cohen and P. P. Davis, Mol. Phys. 33, 819 (1972). J. E. Anderson and W. P. Slichter, J. Chem. Phys. 33, 1922 (1964). J. G. Powles, A. Begum and M. O. Norris, Mol. Phys. 33, 489 (1969). H. A. Resing, Mol. Cryst. Liq. Cryst. 9, 101 (1969). J. E. Tanner, J. Chem. Phys. 39, 3850 (1972). S. Albert and J. A. Ripmeester, J. Chem. Phys. 39, 1069 (1973). P. S. Allen, A. W. K. Khanzada and C. A. Dowell, Mol, Phys. 33, 1273 (1973). I. Perlman, H. Gilboa and A. Ron, J. Magn. Resonance 3, 379 (1972). S. Albert, H. S. Gutowsky and J. A. Ripmeester, J. Chem. Phys. 39, 1332 (1972). 252. 253. 254. 255. 256. 257. 258. 259. 260. 261. 262. 263. 264. 265. 266. 267. 268. 269. '270. 271. 272. 273. 245 J. D. Graham and J. K. Choi, J. Chem. Phys. 93, 2509 (1975). M. J. R. Hoch, J. Chem. Phys. 93, 2522 (1976). J. Virlet and P. Rigny, J. Magn. Resonance 39, 188 (1975). T. Eguchi, G. Soda and H. Chihara, J. Magn. Resonance '33 55 (1976). J. D. Graham and J. S. Darley, J. Magn. Resonance 33, 369 (1976); 33, 287 (1976). G. Janssens, P. van Hecke and L. van Gerven, Chem. Phys. Lett. 33, 445 (1976). L. J. Burnett and B. H. Mu11er, J. Magn. Resonance 33, 343 (1976). L. J. Burnett, J. Chem. Phys. 99, 873 (1977). S. K. Garg, J. Chem. Phys. 99, 2517 (1977). T. Eguchi and H. Chihara, J. Magn. Resonance 39, 409 (1977). R. Kubo and K. Tomita, J. Phys. Soc. Jap. 9, 888 (1954). H. S. Gutowsky and G. E. Pake, J. Chem. Phys. 39, 162 (1950). J. R. Hendrickson and P. J. Bray, J. Magn. Resonance .39, 341 (1973). J. H. Van Vleck, Phys. Rev. 33, 1168 (1948). E. R. Andrew, "Nuclear Magnetic Resonance", Cambridge UP, Cambridge, England (1955). U. Shmueli, M. Polak and M. Sheinblatt, J. Chem. Phys. 39, 4535 (1973). M. Polak, M. Sheinblatt and U. Shmueli, J. Magn. Resonance 39, 252 (1974). H. A. Resing and H. C. Torrey, Phys. Rev. 131, 1102 (1963). D. C. Ailion and C. P. Slichter, Phys. Rev. Lett. 33, 168 (1964). C. P. Slichter and W. C. Holton, Phys. Rev. 122, 1701 (1961). C. P. Slichter and D. C. Ailion, Phys. Rev. A135, 1099 (1964). D. C. Ailion and C. P. Slichter, Phys. Rev. A137, 235 (1965). 274. 275. 276. 277. 278. 279. 280. 281. 282. 283. 284. 285. 286. 287. 288. 289. 290. 291. 292. 293. 294. 246 S. R. Hartmann and E. L. Hahn, Phys. Rev. 128, 2042 (1962). D. C. Ailion, Adv. Mag. Res. 3, 177 (1971). D. C. Look and I. J. Lowe, J. Chem. Phys. 44, 2995 (1965); 3437 (1965). G. P. Jones, Phys. Rev. 148, 332 (1966). J. H. Simons, "Fluorine Chemistry", Academic Press, New York, Vol. 1 (1950). J. G. Aston and T. P. Zolki, J. Am. Chem. Soc. 77, 804 (1955). N. W. Luft, J. Phys. Chem. 39, 92 (1955). E. A. Mason and M. M. Kreevoy, J. Am. Chem. Soc. 33, 5808 (1955). J. 6. Aston, P. E. Wills and T. P. Zolki, J. Am. Chem. S°°°.ZZ 3939 (1955). H. S. Gutowsky, G. G. Belford and P. E. McMahou, J. Chem. Phys. 39, 3353 (1962). E. Catalano and K. S. Pitzer, J. Phys. Chem. 93, 838 (1958). J. Karie, J. Phys. Chem. 33, 4149 (1966). F. B. Brown, J. Mol. Spectrosc. 33, 163 (1967). R. A. Newmark and R. E. Graves, J. Phys. Chem. 33, 4299 (1968). R. A. Pethrick and E. Wyn-Jones, J. Chem. Soc..A54 (1971). Krishnaji and R. K. Laloraya, J. Chem. Phys. 93, 1918 (1974). D. E. Eggers, R. C. Lord and C. W. Wickstrom, J. Mel. Spectrosc. 39, 63 (1976). R. A. Newmark and C. H. Sederholm, J. Chem. Phys. 33, 602 (1965). M. Iwasaki, S. Nagase and R. Kojima, J. Chem. Phys. 22, 959 (1954) — E. R. Andrew and D. P. Turnstall, Proc. Phys. Soc. 93, 986 (1963). I. P. Biryukov, M. G. Voronkov and I. A. Safin, "Tables of Nuclear Quadrupole Resonance Frequencies", Israel Program for Scientific Translations, Jerusalem (1969). 295. 296. 297. 298. 299 O 300. 301. 302. 247 L. Pauling, "The Nature of the Chemical Bond", Third Ed., Cornell UP., New York (1960). F. J. Weigert and J. D. Roberts, J. Am. Chem. Soc. 99, 3577 (1968). W. C. Harris and D. B. Yang, J. Mol. Structure, g, 257 (1973). P. J. Schapiro and J. L. Hoard, J. Am. Chem. Soc. 39, 3347 (1954). J. N. Murrell and Harget, "Semi-empirical SCF 1!) Theory of Molecules", Wiley-Interscience, New York (1972). N. Boden, J. Cohen and R. T. Squires, Mol. Phys. 33, 1813 (1976). C. R. Lassigne and E. J. Wells, J. Magn. Resonance 32, 215 (1977). 13 79 The value J ( C Br)- 30 Hz for CH Br which they report is 3 actually lower than predicted from J (ISCSSCI) and the ratio of gyromagnetic ratios . G. J. Jan: and S. C. Wait, Jr., J. Chem. Phys. _2_6_, 1766 (1957). APPENDIX APPENDIX A. Instructions for NMRLX - A General Program for Compgter Controlled Pulsed NMR Studies I. INTRODUCTION NMRLX is a powerful program used with a Nicolet 1080 series com- puter and a Diablo disk system to do various on-line measurements of nuclear relaxation times and of the self-diffusion constants of molec-. ular motion. In these experiments, the perturbation pulses are trig- gered by software instructions with controllable pulse intervals.’ For single-line spin systems, transient signals are digitized, integrated and analyzed using weighted least-squares fitting. For multiline spin systems the free induction decays are collected and stored in the disk for further data manipulations. This manual describes the structure of this program and the detailed procedures fer each type of experiment. 11. CONSTRUCTION OF THE PROGRAM The program NMRLX is written in assembly language. It occupies totally 4 K memory words each with 20 bits. These 4 K words are. divided into four 1 K-word pages for ease of programming. The whole program can be separated into three main routines: display, pulsing and least squares, plus two embedded auxiliary standard Nicolet pro- grams: Nicolet Debugging Program (N06-30612) and Floating Point Package, 1972 (Nll—20823). The address assignment is shown below: 248 249 0 1112 1663 1777 1 Display 1 Dsd P01nters [ Page 1 2000 2777 3777 1 Discrete Data L Pulsing (T1,T2.T10.Th) l Page 2 4000 4602 S446 S777 Least Squares NICOBUG Subroutines Page 3 l I 6000 7600 7777 FPP-72 Monitor Page 4 I l [ The source tapes are made in pages; each page can be assembled independently. The binary tape of the entire program was also made. It can be read into core by Demon II in the usual way. III. LEGAL COMMANDS The legal commands in NMRLX are mostly composed of two alphanu- meric characters, only a few need one more character as an extension. These commands are divided into three groups: the first group of com- mands accepts parameters which are fixed to integers before being stored; the second group of commands accepts real number parameters which are stored in floating point format; and the third group of commands are executing commands. All the legal commands are listed below: Group 1: NS - N1 - IL - BL - BA - IN - DL — Group II: D" - T3 - 250 Number of scans for each experiment Number of points for each experiment Number of preceding u pulses (used in D5 measurement) d Number of intermediate 1 pulses (used in Dsd measurement) Number of'points used for averaging or maximum finding in data transfer Number of points for integration Number of points fer baseline correction The number to be added to each point Number of points as the interval between the adjacent inten- sified points Number of points delayed before the first intensified points Dwell time Estimated Spin-lattice relaxation time (used in T1 measure- ment) or tau delay in spin-echo pulse train (used in Dsd measurement) Tau delay in Spin-echo pulse train (used in T2 measurement) or time interval between the first field-gradient pulse and the a pulse thereafter (used in D5 d The width of field-gradient pulse (used in Dsd measurement) measurement) Time delay between adjacent pulse sequences 251 Group 111: DS - Display all the data points on the sc0pe DTA - Average the RA points around the intensified points and transfer the average to the discrete data points storage area DTS - Direct transfer of the intensified points to the discrete data points storage area without averaging or searching for maxima ZE - Zero the displayed memory DTM - Search the maximum point within the RA point range around each intensified point and transfer it to the discrete data points storage area DZ - Di5play the current discrete data points on the scope D1 - Display the first discrete data points on the scope' UP - Shift up one block of data points A DN - Shift down one block of the discrete data points DV - Divide the current discrete data points by the previous discrete data points and display the result BC - Use BL points for baseline correction LS - Fit the data points to a straight line using weighted least- squares method EX - Display the pre-analyzed data points AC - Add constant BA to each data point DP - Delete the selected point GO - Start Ds measurement d 61 - Start Tl measurement GZ - Start T2 measurement 252 P1 - Reset page number PRA - Print out the data points in decimal form PRO - Print out the data points in octal form MO — Exit to the monitor NG Go to NICOBUGII £3 Start the triplet or quartet pulse experiment HS '- Set up Homospoil Tl experiment SE Set up mode for pulse adjustment IV. TIMING The basis of timing in this program is the definite executing time for software instructions. (Nicolet 1080 computers have two accurate master crystal oscillators to control the timing.) Since the execu- tion time of each direct instruction is 4 us and is 6'us for indirect instructions (newer models of 1080 computers have 3.3 us and 4.9 us fer direct and indirect instruction, respectively), this timing basis is not for pulse-width-control which is usually only a few us to less than 100 us, except for the spin-locking pulse for T1p measurement. It is only used to control the pulse intervals which are usually longer than a few hundred us. Due to the definite execution time and the limited minimum number of instructions for complex programming, the software-controlled timing has limited accuracy. Theoretically, there is no limit of accuracy for long time delays, but for time delays shorter than 1 ms the accuracy is limited to 1% or less and becomes even greater for a shorter time unless a discrete value is selected. Table I shows these discrete values and the possible errors for arbi- trarily entered values. Tab 1e A1 . 253 Discrete Times and Possible Errors ° the NMRLX Timing Routine Time Entered (us) Actual Time (us) 3 Error 0 N 15 38 153 N m 16 N 31 38 22 N 138 32 N 47 38 0 N 23.4 48 N 63 54 0 N 14.3 64 N 79 70 0 N 11.4 80 N 95 86 0 N 9.5 96 N 111 102 0 N 8.1 112 N 127 118 0 N 7.1 128 N 143 134 0 N 6.3 144 N 159 150 0 N 5.7 160 N 175 166 O N 5.1 176 N 191 182 0 N 4.7_ 192 N 207 198 0 N 4.3 208 N 223 214 0 N 4.0 224 N 239 230 0 N 3.9 240 N 255 246 0 N 3.5 256 N 271 262 0 N 3.3 272.N 287 278 0 N 3.1 288 N 303 294 0 N 3.0 304 N 319 310 0 N 2.8 320 N 335 326 0 N 2.7 336 N 351 342 0 N 2.6 415 N 400 406 0 N 2.2 480 N 495 486 0 N 1.8 254 Generally, when time entered is 16n us to (l6n+15) us, where n is an integer not less than 3, the actual time will be (l6n+6) us and the possible timing error will be 0 to 900/16n+15. Therefore, to avoid serious timing errors one should always choose a value (l6n+6) us unless the time is longer than a few milliseconds. The minimum time which can be provided is 38 us. V. DISPLAY When first started, or after the command has been executed, the program is always in display routine waiting for another command. There are two different types of display; one is for displaying data in memory addresses selected by push buttons on the 1080 mainframe, the other is for displaying discrete data points stored in memory addresses 2000 to 2777. In the first display routine, one can select equally Spaced points to be intensified for transferring to the discrete data storage area.. The intensified points can be selected by three commands DL, NP, and IN. The first intensified point is the (DL+l)th point, the following points to be intensified are separated by IN points and NP is the total number of intensified points. The intensified points can be trans- ferred to the discrete data point storage area by command DT(A,M or S). DTA will average 1/2RA points to the left and the right of each inten- sified point then transfer it to the discrete data point area. DTM will transfer the maximum point in the l/2RA point range to the left and the right of each intensified point. DTS simply transfers the intensified points. This is very useful for picking up spin-echo 255 maxima during the T2 experiment. Setting either DL=0 or NP=0 will abort this function and return to normal continuous display. The second di5play routine displays the data points in storage area 2000-2777 which are stored in the fOrm x1, y1, x2, y2, ... xn, yn. This routine can be terminated by hitting any character on the tele- type and the program will go back to the beginning waiting for a new command. Command D1 will display the first NP point block of the storage area, DZ will display the current NP point block. UP command '4111 shift up one block, DN will shift down one block. DV command is used to divide each point by the corresponding point in the previous block and replace the current block by the result. VI. WEIGHTED LEAST-SQUARES CURVE FITTING A general expression for the decay of magnetization is M(t) = M e‘tfr ; 0 when t gets long M(t) approaches the equilibrium value ”0' On a logarithmic scale small numbers are more seriously scattered and are less important to the fitting. In order to suppress the errors induced by this scattering of small numbers, every data point was weighted by its own amplitude. The mathematical equations for this weighted least- squares curve fitting are shown below: For y1 = -R1Xi + c, where y1 = ln M(tk), Xi = ti, R1 = relaxation rate = 1/T1 "’256 I . 2 Z"i inyiwi ' inwf Z’iwi l/Tl 3 R1 = 2 2 2 2_ s (1) (inwi) — {wi {xiwi M(t.) 1 C where Wi - W , 8.150, . 1 1 2 2 2 22 . 2x1": inyiwi ’ Zyiwi inwi c 2 2 2 2 2 (2) (zxiwi) ' 2"i {xiwi and the standard deviation, P x. '- chl-+ c - v.12 1’2 2 1 1 (3) o = , Xnax-den L N - l d where N is the number of data points used for curve fitting. When discrete data points have been collected and displayed on the scope, they have to be baseline corrected in order to make every point positive in magnitude. Baseline correction is via command BC which will use the last BL points as the baseline. Command LS starts the least- squares calculation and then prints out the slope and l/slope of the theoretical line and the corresponding standard deviations. The scope will show the deviation of each point from the theoretical line. Data can be recalled by command EX and can have a constant BA added by com- mand AC to shift the baseline up or down. The best fit can be seen from the minimum standard deviation printed out and the uniform dis- tribution of the deviations from the line as seen on the scOpe. 257 VII. SETUPS FOR EXPERIMENTS (1) Pulse Adjustment Before starting data collection, the pulse width has to be adjusted to the desired width. This can be done by using setup mode (SE) which can trigger an even number (determined by RA) of equivalent pulses with interval controlled by N1. Normally when adjusting a n/Z or n pulse, RA is set to be 4n (n is any integer) and N1 is set to be (pulse interval/l6). The pulse inter- val has to be long enough to see the signal and short enough to avoid saturation of the signal after a pulse sequence. TL (delay between each pulse sequence) has to be set long enough to avoid an excessively heavy duty cycle for the power amplifier. When the pulse width is adjusted to be exactly 90°, the signal on the scape should be up-null-down-null-etc. The signal will be nulled when the pulse width is exactly 180° and also when the RF phase is adjusted to be 90° (along the Y axis). If the RF pulses are not homogeneous, there will be some deviation from the uniform up-null- down-null pattern. Therefore, this procedure also offers.a method for testing RF pulse homogeneity. This setup mode will always trigger pulse 1 of the computer. If one needs to adjust the second or the third pulse, one can either con- nect the pulse 1 cable to the second or third pulse input or use NICOBUGII (called by command NC) to change address 3745/4102 to 4104 or 4106. 258 INA") an) 'H\§r n * 00,021) vo.(P1) b I 'k T L = FG P K,(P1) P2 C T L 4 S L P A we. . 00y . no, my 6 l /f\\; W/:\\ I L] . TL ‘ ' : I 5 //\\ . 2 ,1, I noy soIt no, 00.. e p '1 I; I; FGP [PCT] F— T L ——_" no, 90,I looy noy 90., so, . no, 00,, 1 ‘ h 3) 2‘) 33 «— TL —. FGP FGP nay ”a '”y Qxfi'w ”.100. no, 00_ ”'0 no eq- 9 I: I; I: I: I; *— TL ——9 FG P FGP .0. I'D, 100,342.— "1001”,”, 'I0, IIOy h .- I)‘ b '1'. I. <— T L _. IFGPL; . [FGP j\ 877 .44 * A Figure 1. Pulse sequences used in NMRLX. 259 After the pulses have been well adjusted, one can start to set up parameters for each type of experiment. Pulse sequences for these experiments are shown in Fig. 1. Typical procedures are given in the following sections. (ii) T1 Measurement Using the Spin-Inversion-Recovery Techniqge a. Adjust PULSE A to be (n,x) and PULSE B to be (n/2,-x). (x,y,z are the laboratory coordinates, z is in the direction of HO, the static magnetic field,and y is the direction of the coil axis.) b. For a single-line spin system set on resonance, set: T1 = Estimated spin-lattice relaxation time of the spin to be measured NP Number of points exploring the exponential decay of magnetization, therefore number of r delays NS = Number of scans to be averaged for each point TL At least 5 times the estimated T1 IL Number of points to be integrated as the amplitude of magnetization for each I delay. It can be from 1 to the size of the di5played memory selected by 1080 mainframe buttons. Then type 61 and answer "N" for the question FOR MULTILINE? immediately after 61. The computer will use the T1 value entered to calculate r 4T 1 delays Ti = ljfiy-, i = l to NP, then trigger the pulses 180—ri-90 and collect the transient signal (usually 20 us dwell time and 1 K memory size are used), integrate it and store the amplitude in storage area 2000-2777. Instead of collecting NS scans with each Ti then inte- grating the averaged signal, the program will increment r after each 260 scan and collect NP points first then do another scan and add the result to the previous cumulated amplitudes stored. The advantage in this procedure is that it minimizes the error induced by field drift during long time averaging. Of course, this can also be achieved by using a smaller IL. After the experiment is done the scope will dis- play NP points with the x axis representing 1 delays and the y axis representing amplitudes. The least-squares routine can then be used to fit the exponential decay to a straight line and print out slope (R1), l/slope (T1) and the standard deviations. c. For multiline spin systems only T1, NP, NS, and TL are needed. Answer "Y" for the question FOR MULTILINE? after which Gl will begin data collection. The computer will first check if tracks 400-625 have any program or data stored. If that is true, an error message NO MORE ROOM will be printed out and stop any action thereafter. If not, the computer will zero those tracks then start collecting FID's corre- sponding to each 1 delay and stored them adjacently on the disk starting at.xrack 400. The memory size and dwell time should be set on the 1080 mainframe before data collection. After the experiment is done the scope will show the cumulated FID of the last r delay for fur- ther Fourier transform using the auxiliary program FTSWAP (see description in Section X). (iii) T1 Measurements Usigg the Homospoil Techniqge a. Adjust PULSE A to be (n/2,-x) and connect PULSE C to field- gradient pulse input. b. Enter parameters as in (i). 261 c. Type command HS then start data collection by 61. d. The analysis is the same as in (i). (iv) T10 Measurement a. Adjust PULSE A to be (n/2,x) and PULSE B to be (n/2,y). b. Push in buttons marked Tlp so that PULSE 2 becomes the ter- minating trigger of the spin-locking pulse right after PULSE 1. Note: An occasional circuit hangup might turn the -15 v pulse to the power amplifier always on which can be noticed on the current meter of the 500 V plate power supply. If this occurs release the T19 button then push it in again. c. Enter parameters Tl, NP, NS, IL (for single-line system) and TL as before and use the command 01 to start data collection. d. Data analysis is the same as in (i). (v) T2 Measurement a. Adjust PULSE A to be (n/2,x) and PULSE B to be (n,y). (Steps b-e are for a single-line system, f-h are for a multiline sys- tem.) b. For a single-line system set on resonance, connect cable T2 to SWEEP CONTROL CONNECTOR so that pins X, H, F are plugged in. The other end of the cable should be connected to a pulse inverter then to the pulse B input of the pulse-width controller. c. Enter T2 equal to the r delay in the pulse sequence, DW equal to the number selected on the 1080 mainframe and set ADDRESS SELECTOR 262 to a number n which fulfills the relation T2 = DW-n and DELAY MODE to OFF. In this case, when the MEASURE mode is in process there will be a (n,y) pulse triggered for every n addresses advanced. d. Enter NS to be the number of scans and TL to be at least 5T1, then use command 62 to start data collection after answering "N" for the question FOR MULTILINE? e. When the experiment is done the scope will show decaying spin echoes; the display routine can then be called to pick up the echo maxima and transfer them to the discrete data point storage area for least-squares analysis. f. For a multiline system disconnect cable T2 and set the memory size and dwell time on the 1080 mainframe as in the FTNMR experiment. g. Enter T2 (1 delay), NP, NS, and TL as usual and answer "Y" for the question FOR MULTILINE? after the GI command. h. The analysis is the same as for multiline T1 measurements. Note: In order to decrease errors from pulse misadjustment, only odd echo FID's are collected. (vi) Th Measurement Th bination of T1 and T2. For pulse sequences e-q in Fig. 1 the apparent is defined as the hybrid relaxation time and is a linear com- relaxation time can be expressed by Th'l = (I-p) Tl-l + pTz-1 . (1) 263 where p = th/t1+2t2 (or p = 4t2/t1+4t2 for pulse sequences f or g) and the decay of the signals follows the expressions Mt - Moo = (MO-Moo)exp(-t/Th) ,’ (2) where Mo is the magnetization at t = 0 and Mm is related to M6 by 1-8XPI-t1/71] M = M m o l-eXPl-(t1+2t2)/Th] ° ‘33 If pulse sequences f or g of Fig. l are used, 4t2 replaces th in Equation (3). The principal advantage of this technique is a great saving in time and it should therefore be applicable to the measurement of T1_and T2 in natural abundance 13C or 15N. The following procedure is used to set up the Th experiment: 3. Adjust PULSE A to be (n,y) and PULSE B to be (n/2,x). “ 264 b. Connect the phase inverter as follows: J rModified Part . I , Phase Invertor r ---------------------- ‘n PULSE Controller : To PC PULSE C Current V Mixin to Pulser RF Gaté—4pto Mixer Mixer - Network 5 PA Gate .4pto PA c. Enter the following parameters: T1 = Estimated Th DW = Hardware setting oanWELL TIME on the 1080 front panel N1 = Hardware setting of DELAY TIME on the 1080 front panel NP = Number of FID's to be collected NS = Number of scans d. Type the TR command to start data collection. Note 1: Three different pulse sequences may be chosen. These are determined by address 3601. When it contains 4104, e sequence will be chosen; when it contains 1676, f sequence will be chosen; and when it contains 1673, g sequence will be chosen. They can be selected by using NICOBUGII. 26S 4T1 NP routine ZRMEM and ADD. Therefore t1 = { Note 2: t1 in Eq. (1) is plus the computer time for executing sub- 4T1 NP 34(SIZE/1024 - 1)} , where SIZE is the memory size used for collecting + 240 us + 96 x SIZE + each FID; t2 in Eq. (I) is equal to [N1 + DW x SIZE]. e. Data analysis is the same as in (i) for multiline system. (vii) DSd Measurement The pulsed field-gradient spin-echo technique (pulse sequence h in Fig. l) is used in this program to measure Dsd' The echo amplitudes with the field-gradient pulse (FGP) on and off follow the relation 1n §é(%-= -v2g26205d(A-6/3) . (4) where M(t) is the echo amplitude with PCP on, M°(t) is the echo ampli- tude with PCP off, g is the magnitude of the field gradient, 6 is the duration of PCP and A is the diffusion time. To set up this experiment follow the procedures below: a. Connect PULSE C to PG CURRENT PULSER trigger IN. Turn the power on. b. Enter parameters T1, T2, T3 (see Fig. l.h) N1, N2, NP, N, IL, and TL. Note: T1>T2>T3 and 2NP§(N2+1) have to be fulfilled. c. Adjust FGP duration and FG current by doing a dummy run (GD command), monitoring the PCP duration from the BNC connector and the current from the pin test-point on the front panel of PG CURRENT PULSER (+3 V corresponding to ~10 amps current). 266~ d. Turn the PCP trigger control switch off and run once (without FGP) by typing GO command. Note: A message "Tl,T2&T3 MUST BE <8 SEC" will be printed out immediately after GO command. This was done to save memory in pro- gramming. Actually no such experiment needs T1 longer than 8 seconds. e. Turn the PCP trigger control switch on and run once again. (After each run the scope should display data points collected.) -f. Type DV command to display the result of M(t)/M°(t). g. Call the LS routine to analyze the data. The slope will be equal to yzgzézDsd. Therefore, Dsd = slope/yzgzoz. Note: Due to insufficient heat insulation of the field-gradient coil, the actual field gradient is temperature dependent. A calibration has been done using CFC13 for which accurate Dsd values have been reported. VIII. USEFUL SUBROUTINES In NMRLX, several subroutines are used repeatedly. They are con- structed rather independently and can be called for use at any time. 1. TIMER: This subroutine is actually a counting clock. The basic logic of its construction is to count down a number stored pre- viously in address WAIT. One count loop takes 4 instructions and there- fore 16 us (or 13.2 us for newer models of 1080, see Nicolet 1080 manual for instruction execution time). Since NIC 1080 has 20 bits in one word the maximum counting time for each word is about 8 seconds. Therefore, another address HIWAIT is used to store the number of times of counting down the number stored in WAIT. 267 2. TMS: This subroutine is used to set up WAIT and HIWAIT. It must be called before using the TIMER subroutine. The time entered in seconds should be put in FAC of the Nicolet FPP-72 package before calling TMS. A typical set up for a timer is: JMS @GETAC TI IVE JMS TMS JMS TIMER Note: A floating-point constant CONFPI in addresses 1105 and 1106 is used in subroutine TMS to calculate WAIT. For models with 4 us exe- cution time it is 105/l6 (1105/4o,ooo; 1106/1,720,440). For those with 3.3 us execution time it should be changed to 106/13.2 (1105/ 42,466; 1106/1,117,676). 3. ZRMEM: A.subroutine to zero the displaying memory addresses. 4. ZRTK: A subroutine to zero the content on chosen tracks in the disk. The first and the last track numbers to be erased should be put into addresses TRACKI and TKEND. 5. ADD: A subroutine to add displaying memory content to the content stored in selected tracks. Ix. ERROR MESSAGES NO MORE ROOM: Discrete data storage area is full or tracks on the disk have been used. RA T00 BIG: Assigned number of points in RA exceeds the dis- playing memory size. BL TOO BIG: Number of points assigned for baseline correction exceeds total discrete data points displayed. 268 A,M OR 8 MUST FOLLOW: Other extention is used after DT. UNEQUAL POINTS: Number of x's is not equal to the number of y's. ?? SET N2 = ODD, RA = EVEN: Even number (or odd number) has been selected for N2 (or RA). . 0 OR A MUST FOLLOW: Other extention is used after PR. T1<=T2 MUST FOLLOW: Obvious. T2<=T3 MUST FOLLOW: Obvious. 2NP>N2+1 MUST FOLLOW: Obvious. X. AUXILIARY PROGRAMS l. FPP72: Standard Nicolet Floating-Point Package, 1972 version (see Nicolet manual). 2. NICOBUG: Standard Nicolet Debugging Program (see Nicolet manual). 3. FTSWAP: A modified Nicolet FTNMRII for fetching FID's stored on the disk, doing the Fourier transforms and storing spectra back on the disk. Command to be used: NT: Number of tracks for each FID (l for l K words, 2 for 2 K, 3 for 4 K and 6 for 8 K). ST: Starting track nunber wherethe FID's were previously stored. (It is preset to 100,400 (in octal) track number 400.) NS: Number of FID's stored and to be transformed. NM: Select a three-character file name for the spectra to be stored. Series numbers from 001 to nnn will be added automatically when spectra are stored. '269 L1: Link a number of normal FTNMRII commands needed for doing data processing, e.g., TR, EM, FT, PC, TR, PP, etc. 80: Start execution of fetching the FID's, processing of data according to commands linked by LI, then storing spectra onto the disk with file name selected by NM. XI. USEFUL PATCHES Since a large variety of pulse NMR techniques have been developed., it is impossible to include everything in a single program. Therefbre, a few small patches have been written in modular form which can be read into NMRLX to do other experiments. 1. SSFT: A.patch to do single-scan FT multiline T1 measurement with pulse sequence 180-1-30°-FID-r-3O°-FID-r...1 The pulse sequence and the decay of magnetization are illustrated below. 180 30 so 30 30 30 L7 igEaLT 7&1 7 T 17:61 . [a [E]. F r f 2. STR: A patch to do a single-line Th experiment. For single- line spin systems there is no need to use FT. Therefore, t2 in the pulse sequence Fig. le-g can be set short enough to ignore spin- spin relaxation effects and the transient signal can be integrated as in the single-line T1 experiment. The resulting decay constant will no longer be Th’ it will simply be T1. This patch occupies 3034 to 3131. 270 3. DYNOE: A patch to measure the dynamic Overhauser effect. It utilizes the homospoiling T1 measurement module of NMRLX and has only two addresses modified: 3773/111760 changed to 111762 3762/4106 changed to 4104. A decoupler gating box should be connected as follows: PULSE A Pulse width to normal trans- controller . mission line to receiver PULSE B 1V - line Decoupler [P jFlReceiver PULSE C Gatin Box . 7 , :l g . m Gate , _ Signal Decoupler to Decoupler RF Source Coil PULSE B is then used to trigger the decoupler and PULSE C to turn off the decoupler. The pulse sequence is: PULSE B A C B A C Pulse . Sequence 7 TL 7' ~_— or g 2:: ON[] Jill] :[I {1 1L*- [n.n.uuu..n..n.] I, Decoupling RF "'°3:_flfl_ HI ,, I ...................... [— 271 XII. INTERFACING CIRCUITS 1. Pulse booster: Convert PULSE 1 and PULSE 2 from NIC 1080 (500 nsec) to PULSE A, PULSE B and PULSE C (O to S V, 2‘Usec, SmA) (see Reference 131 for description). 2. Pulse width controller: Convert two 0 to 5 V pulses to two 0 to ~15 V pulses with pulse width adjustable from 1 to 100 psec for driving the NMRS MP-lOOO spectrometer (see Reference 131 for description). 3. FG current pulser 1: Supply an externally triggered field- gradient pulse with lOlisec to 10 msec pulse width and 0 to 10 A current to a field-gradient coil built in the NMR probe (see Reference 133 for description). 4. FG current pulser 2: Supply an externally tiggered field- gradient pulse with l to 100 msec long and 25 mA current to the z shim coils of the magnet (see Reference 131 for description). 5. rf phase inverter: Reverse the output rf phase everytime it is triggered (see Reference 133 for description). 6. Decoupler gating box: Provide gating pulses for decoupling rf to the power amplifier and for the receiver. (see Figure A2). 272 . 8 5 0.3 «3:333.— 28 a: 5 can moi—«onus u .Enuuumv 33.39 x2— ufiuuu .3330qu .2 can»: ..N :5 E : 3 009 xu 0+ 273 INSTRUCTIONS FOR OTHER PROGRAMS USED 1. FTNMRD - A modified Nicolet FTNMRII for triggering a pulse to do FTNMR. Additional commands: NS - Number of sweeps DL - Delay in seconds after each pulse and data collection GO - Starts triggering perturbation pulse and collects FID. TOTAL EXPERIMENT TIME will be printed out. CO - Resume data collection after interruption II. FTNMRB - A modified Nicolet FTNMRII for collecting data after a pulse triggered by NIC-293 general pulser. Additional comands: P2 - Enter pulse width in “sec DZ - Enter delay time in seconds after each pulse and data col- lection NC - Number of scans GO - Start data collection after total experiment time is printed out CO - Resume data collection after it has been interrupted 274 III.SECMFT - This is a modified FTNMRD or FTNMRB for calculating the second moment of the selected peak. Additional legal commands: GM - Gyromagnetic ratio divided by 2n M2 - Enter second moment calculation mode and intensify the peak selected Subcommands of M2 mode: N - Calculate and print out unnormalized second moment C - Calculate and print out normalization factor G - Calculate and print out normalized second moment Procedure for the experiment: 1. Use normal FTNMRII functions to collect data and transform the FID to a spectrum. 2. Select F1 and F2 to confine the peak whose second moment is to be calculated. 3. Set cursor to the center of the peak. 4. Enter correct GM for nucleus (4259.1 for 1H and 4007.4 1F). 5. Type M2 to intensify the peak area and then type N, D and finally G. The second moment of the right and left sides of the peak will be printed out. (This is designed for separating the contribu- tions from an embedded peak which causes asymmetry of the overall peak.) Note: In order to get a correct second moment the baseline of the selected peak has to be correctly set. Command AC of FTNMRII is very useful for doing this.