ALKALIMETAL‘NMRSTUDIESOFSODIUM,=RUBIDIUM ¥ _ _ AND CESI‘UM, ANIONS AND THE KINETICS 0F SODIUM «_ _ , l ' CRYPTATE EXCHANGE REACTIONS ’ . Dissertation for the Degree of PH, D.’ MICHIGAN STATE UNIVERSITY . I I N‘JOSEPHIMIICHAELCERASO - 1 . 1975 , ' ' Date 0-7639 This is to certify that the thesis entitled ALKALI METAL NMR STUDIES OF SODIUM, RUBIDIUM AND CESIUM ANIONS AND THE KINETICS OF SODIUM—CRYPTATE EXCHANGE REACTIONS presented by Joseph Michael Ceraso has been accepted towards fulfillment of the requirements for Pho Do dpgfpp in Chemstry Major professor 9-29-75 ABSTRACT ALKALI METAL NMR STUDIES OF SODIUM, RUBIDIUM AND CESIUM ANIONS AND THE KINETICS OF SODIUM-CRYPTATE EXCHANGE REACTIONS By Joseph Michael Ceraso The solvent influence on the kinetics of the release of sodium ion from its complex with the macrobicyclic hexaoxa- diamine N(CH2CH20CH2CH20CH2CH2)3N (2,2,2 cryptand) has been investigated by using the pulsed 23Na Fourier transform NMR technique. Each of the four solvents used, water, ethylene- diamine (EDA), pyridine (PYR), and tetrahydrofuran (THF), showed two well-defined resonance absorptions below the corresponding coalescence temperature. In all cases it was possible to separately measure line widths and chemical shifts of the free and complexed sodium ion as a function of temperature in the absence of exchange. Since the spins are non—interacting, the transient solution of the modified Bloch equations which describe the response of a spin system under— going chemical exchange between two nonequivalent sites is applicable. Exchange times, T, were calculated by fitting Joseph Michael Ceraso the modified Bloch equations to the observed 23 Na line shape by using a weighted non-linear least-squares technique. For EDA solutions, a mechanism in which the rate limiting step is release of sodium ion from the cryptate complex was found to be in agreement with the observed concentration dependence of exchange times throughout the entire temperature range examined. The activation energy for sodium cryptate exchange was found to vary from 12.9 kilocalories for EDA solution to 16.7 kilocalories for H20 solution. When 2,2,2 cryptand and an excess of sodium, rubidium or cesium metal are placed into contact with THF, ethylamine or methylamine, solutions at least as concentrated as 0.1 M total metal were obtained. Major species in solution are the complexed cation and the alkali metal anion M-. The NMR chemical shift and linewidth have been measured for 23Na- in THF, EA and MA, for 87Rb- in THF and EA and for 133Cs- in THF. The chemical shift of Na- is, within experimental error, the same as that calculated for the gaseous anion (based upon the measured value for the gaseous atom) and is independent of solvent. Comparison with the solvent-dependent chemical shift of Na+ provides conclusive evidence that Na- is a "genuine" anion with two electrons in a §§_orbital which shield the gp_electrons from the influence of solvent. The linewidth increases from THF to EA to MA, suggesting either an increasing exchange rate with the cryptated cation or, more probably, the influence of an increasing concentration Joseph Michael Ceraso of solvated electrons. In the case of sodium solutions in all solvents both Na+C222 and Na' are detected by their NMR peaks. However, probably because of extreme line-broadening, Rb+C222 and Cs+C222 are not observed, but only the relatively narrow line of the corresponding anion. The chemical shifts (dia- magnetic shift in ppm from the infinitely dilute aqueous ion) are 185 and 197 for Rb_ in BA and THF, respectively, and 292 for Cs‘ in THF compared with 212 and 344, respectively, for the gaseous Rb and Cs atoms. A synthesis of the first crystalline salt of the sodium anion is described. The salt has the stoichiometry Na+C Na- 222' (where C222 is 2,2,2 cryptand) and is gold in color at low temperatures. ALKALI METAL NMR STUDIES OF SODIUM, RUBIDIUM AND CESIUM ANIONS AND THE KINETICS OF SODIUM-CRYPTATE EXCHANGE REACTIONS BY Joseph Michael Ceraso A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1975 To my Parents ii ACKNOWLEDGMENTS The author wishes to express his sincere appreciation to Professor James L. Dye for his endless encouragement and guidence throughout this research program. Special thanks go to Steve Landers and Patrick Smith for their collaborative efforts in the exchange studies. Thanks go to Professor Max T. Rogers for being second reader and to Drs. Yves Cahen, Mei—Tak Lok and F. J. Tehan for assistance in synthesizing 2,2,2 cryptand. Additional thanks go to C. W. Andrews, Dr. M. G. DeBacker, Miss E. Mei, Dr. L. D. Long, Dr. N. Papadakis and Dr. D. A. Wright for their help and suggestions. Financial assistance from the Atomic Energy Commission and Michigan State University and a summer term of fellow- ship from the Dow Chemical Company is acknowledged. iii TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . LIST OF FIGURES. . . . . . . . . . . . . . . . . . . . . I. INTRODUCTION . . . . . . . . . . . . . . . . . II. HISTORICAL. . . . . . . . . . . . . . . . . . . . 2.1 Metal Ammonia Solutions. . . . . . . . . . . 2.1.1. Species in M-NH3 Solutions . . . . . 2.1.1.1. Evidence for Monomer Species in M—NH3 Solutions . . . . . . . . 2.1.1.2. Evidence for Diamagnetic Species in M-NH3 Solutions. . . . . . 2.1.2. Models for Metal Ammonia Solutions . . 2.2 Alkali Metals in Amines and Ethers . . . . . . 2.2.1. General Properties . . . . . . . . 2.2.2. Major Species Involved in Metal—Amine and Metal-Ether Solutions. . . . . . 2.2.2.1. Solvated Electrons. . . . . . . 2.2.2.2. Monomers. . . . . . . . . . . . 2.2.2.3. Spin-paired Species . . . . . . 2.2.2.4. Alkali Metal Anions . . . . . 2.3 Overall Equilibrium Scheme . . . . . . . III. GENERAL FEATURES OF ALKALI METAL ION NMR . . . . 3.1 Introduction . . . . . . . . . . . . . . . 3.2 Shielding Constants of Alkali Metal Ions iv Page viii xi 10 12 15 15 16 l6 18 19 20 26 28 28 31 Chapter Page 3.2.1. Alkali Metal Ions in Crystals. . . . . 32 3.2.2. Alkali Metal Ions in Solution. . . . . 38 3.3 Quadrupole Relaxation of Alkali Metal Ions in Solution . . . . . . . . . . . . . . . . . . . 42 IV. EXPERIMENTAL. . . . . . . . . . . . . . . . . . . . 47 4.1 General Techniques . . . . . . . . . . . . . . 47 4.1.1. Glassware Cleaning . . . . . . . . . . 47 4.1.2. Vacuum Techniques. . . . . . . . . . . 47 4.2 Metal Purification . . . . . . . . . . . . . . 48 4.2.1. Storage of Alkali Metal in Small Quantities . . . . . . . . . . . . . . 48 4.2.1.1. Sodium Metal. . . . . . . . . . 48 4.2.1.2. Cesium and Rubidium Metals. . . 48 4.3 Solvent Purification . . . . . . . . . . . . . 50 4.4 Preparation and Purification of 2,2,2 Cryptand . . . . . . . . . . . . . . . . . . . 52 4.4.1. Synthesis of Diethyl Ester of Triglycolic Acid . . . . . . . . . . . 53 4.4.2. Hydrolysis of the Diester. . . . . . . 54 4.4.3. Purification of 2,2,2 Cryptand . . . . 54 4.5 Synthesis of a Crystalline Salt of the Sodium Anion. . . . . . . . . . . . . . . . . . . . . 55 4.6 Solution Preparation . . . . . . . . . . . . . 58 4.6.1. Salt Solutions for NMR Exchange Studies. . . . . . . . . . . . . . . . 58 Chapter Page 4.6.2. Preparation of Metal Solutions . . . . 58 4.7 The NMR Spectrometer . . . . . . . . . . . . . 61 4.8 Temperature Control and Calibration. . . . . . 63 4.9 Data Reduction . . . . . . . . . . . . . . . . 64 V. SODIUM-23 NMR STUDY OF SODIUM ION - SODIUM CRYPTATE EXCHANGE RATES IN VARIOUS SOLVENTS . . . . . . . . . 66 5.1 Introduction . . . . . . . . . . . . . . . . . 66 5.2 Determination and Interpretation of the Line Shapes . . . . . . . . . . . . . . . . . . . . 68 5.2.1. Measurements in the Absence of Exchange . . . . . . . . . . . . . . . 70 5.2.2. Evaluation of Exchange Times . . . . . 88 5.2.2.1. Exchange Times in EDA Solutions . . . . . . . . . . . 90 5.2.2.2. Exchange Times in THF, H20 and PYR Solutions . . . . . . . . . 94 5.3 Results and Discussion . . . . . . . . . . . . 106 5.3.1. Some Sources of Systematic Error . . . 106 5.3.2. Mechanism of Exchange. . . . . . . . . 121 VI. ALKALI METAL NMR STUDIES OF SODIUM, RUBIDIUM AND CESIUM ANIONS . . . . . . . . . . . . . . . . . . . 131 6.1 Introduction . . . . . . . . . . . . . . . . . 131 6.2 Magnetic Shielding Constants of Alkali Metal Ions . . . . . . . . . . . . . . . . . . . . . 131 6.3 Results. . . . . . . . . . . . . . . . . . . . 133 vi Chapter Page 6.4 Discussion . . . . . . . . . . . . . . . . . 134 APPENDIX A - MODIFICATION OF RELAX 2 . . . . . . . . . . 143 APPENDIX B - PROGRAM CONVERT . . . . . . . . . . . . . . 145 APPENDIX C - SOLUTION TO MODIFIED BLOCH EQUATIONS. . . 146 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . 153 vii Table II III IV VI VII VIII IX LIST OF TABLES Page Values of the average excitation energy A, and the expectation value p for alkali metals. . . 37 Experimental and calculated chemical shifts for rubidium and cesium ions in halide crystals and aqueous solution. . . . . . . . . . . . . . . . . . 39 Chemical shifts of free and cryptated sodium ions in solution . . . . . . . . . . . . . . . . . . . . 41 Temperature calibration for NMR . . . . . . . . . . 65 Variation of the relaxation rate and chemical shift of free sodium cation with temperature in ethylenediamine . . . . . . . . . . . . . . . . . . 75 Variation of the relaxation rate and chemical shift of bound sodium cation with temperature in ethylenediamine . . . . . . . . . . . . . . . . . . 76 Variation of relaxation rate and chemical shift of free sodium cation with temperature in water. . . . 77 Variation of relaxation rate and chemical shift of bound sodium cation with temperature in water . . . 78 Variation of the relaxation rate (1/T2) and chemical shift of free sodium cation with temper- ature in tetrahydrofuran. . . . . . . . . . . . . . 79 viii Table XI XII XIII XIV XV XVI XVII Variation of relaxation rate and chemical shift of bound sodium cation with temperature in tetrahydrofuran . . . . . . . . . . . . . . . . . Variation of relaxation rate and chemical shift of free sodium cation with temperature in pyridine. . . . . . . . . . . . . . . . . . . . . Variation of relaxation rate and chemical shift of bound sodium cation with temperature in pyridine. O O O O O O O O O O C O O O O O O O O 0 Activation energy, Er’ for solvent reorganization in various solvents . . . . . . . . . . . . . . . Temperature dependence of the exchange time in EDAl and corresponding relaxation rates in the absence of exchange . . . . . . . . . . . . . . . Temperature dependence of the exchange time in EDA2 and corresponding relaxation rates in the absence of exchange . . . . . . . . . . . . . . . Temperature dependence of the exchange time in EDA3 and corresponding relaxation rates in the absence of exchange . . . . . . . . . . . . . . . Temperature dependence of the exchange time in H20 and corresponding values for relaxation rates in the absence of exchange. . . . . . . . . . . . ix Page 80 81 82 89 101 102 103 111 Table XVIII XIX XX XXI XXII Page Temperature dependence of the exchange time in PYR and corresponding relaxation rates in the absence of exchange . . . . . . . . . . . . . . . 112 Temperature dependence of the exchange time in THF and corresponding relaxation rates in the absence of exchange . . . . . . . . . . . . . . . 113 Comparison of exchange times with and without intensity correction for EDA3 . . . . . . . . . . 120 Exchange rates and thermodynamic parameters of sodium cryptate exchange in various solvents. . . 129 Selected list of shielding constants and linewidths. . . . . . . . . . . . . . . . . . . . 136 LIST OF FIGURES Figure Page 1 2,2,2 Cryptand and l8-crown—6. . . . . . . . . . . 3 solv (1), Cs_ (2), K‘ (3) and 2 Spectra at 250C of e Na_ (4) in THF in the presence of cryptand or crown. . . . . . . . . . . . . . . . . . . . . . . 21 3 The relation between the peak position of Na' and of K_ in diisopropyl ether (DIPE), diethyl ether (DE), hexamethyl phosphoric triamide (HMPA), tetrahydrofuran (THF), dimethoxyethane (DME), diglyme, ethylamine (EA), 1,2—propanediamine (PDA) and ethylenediamine (EDA) at 250C. . . . . . . . . 23 4 The relation between the peak position of K‘ and -— . . O of esolv in various solvents at 25 C . . . . . . . 24 5 (a) Effect of quadrupole coupling upon nuclear Zeeman energies in first order. (b) Hypothetical spectrum corresponding to energy levels of (a) . . 3O 6 Semilog plot of 1/T2 for 23Na vs the reciprocal absolute temperature for a solution containing 0.15 M NaBr in EDA . . . . . . . . . . . . . . . . 45 7 Apparatus for the preparation of storage tubes for sodium metal . . . . . . . . . . . . . . . . . 49 8 Apparatus for preparation of small ampoules of rubidium and cesium metal. . . . . . . . . . . . . 51 9 Vessel for preparation of Na+C - Na- . . . . . . . 56 xi Figure Page 10 Vessel for preparation of NMR samples. . . . . . . 59 11 Block diagram of the multinuclear magnetic resonance spectrometer . . . . . . . . . . . . . . 62 12 A typical KINFIT analysis of the 23Na lineshape for a solution containing 0.60 M NaBr in EDA. X represents an experimental point, 0, a calculated point, =, an experimental and calculated point which are the same within the resolution of the plot . . 74 13 Semilog plots of l/T2 for 23Na vs reciprocal absolute temperature for solutions containing 0.6 M NaBr (o) and 0.3 M Na+C222 - Br‘ (0) in EDA. 83 14 Semilog plots of l/T2 for 23Na vs reciprocal absolute temperature for solutions containing 0.2 M NaCl (0) and 0.2 M Na+C222 - c1— (0) in H20. 84 15 Semilog plots of 1/T2 for 23Na vs reciprocal absolute temperature for solutions containing 0.4 M Na¢4B (o) and 0.2 M Na+C222 - ¢4B‘ (o) in THF . . . . . . . . . . . . . . . . . . . . . . 85 16 Semilog plots of 1/T2 for 23Na vs reciprocal absolute temperature for solutions containing 0.2 M Na¢4B (o) and 0.2 M Na+c222 ° ¢4B“ (o) in PYR . . . . . . . . . . . . . . . . . . . . . . 86 17 Spectra at various temperatures for a solution of 0.15 M C222 and 0.6 M NaBr in EDA (1 ppm 2 15.87 Hz). . . . . . . . . . . . . . . . . . . . . 91 xii Figure Page 18 Spectra at various temperatures for a solution of 0.30 M C222 and 0.6 M NaBr in EDA (1 ppm = 15.87 Hz). . . . . . . . . . . . . . . . . . . . . 92 19 Spectra at various temperatures for a solution of 0.45 M C222 and 0.6 M NaBr in EDA (1 ppm = 15.87 Hz). . . . . . . . . . . . . . . . . . . . . 93 20 Computer fit of spectra obtained with 0.15 M C222 and 0.6 M NaBr in EDA. (a) 42.6OC; (b) 25.50c. . . . . . . . . . . . . . . . . . . . 95 21 Computer fit of spectra obtained with 0.15 M C222 and 0.6 M NaBr in EDA. (a) 64.9OC; (b) 53.90c. . . . . . . . . . . . . . . . . . . . 96 22 Computer fit of spectra obtained with 0.30 M C222 and 0.6 M NaBr in EDA. (a) 45.00c; (b) 30.4OC. . . . . . . . . . . . . . . . . . . . 97 23 Computer fit of spectra obtained with 0.30 M C222 and 0.6 M NaBr in EDA. (a) 76.4OC; (b) 50.20c. . . . . . . . . . . . . . . . . . . . 98 24 Computer fit of spectra obtained with 0.45 M c222 and 0.6 M NaBr in EDA. (a) 30.6OC; (b) 19.3Oc. . . . . . . . . . . . . . . . . . . . 99 25 Computer fit of spectra obtained with 0.45 M C222 and 0.6 M NaBr in EDA. (a) 41.9OC; (b) 36.10c. . . . . . . . . . . . . . . . . . . . 100 xiii Figure 26 27 28 29 30 31 32 Page Computer fit of spectra obtained with 0.2 M c222 and 0.4 M NaI in H20. (a) 23.60c; (b) 3.30c . . . . . . . . . . . . . . . . . . . . 104 Computer fit of spectra obtained with 0.2 M c222 and 0.4 M NaI in H20. (a) 39.30c; (b) 26.9OC. . . . . . . . . . . . . . . . . . . . 105 Computer fit of spectra;obtained with 0.2 M C222 and 0.4 M Na¢4B in PYR. (a) 93.2Oc; (b) 66.7OC. . . . . . . . . . . . . . . . . . . . 107 Computer fit of spectra obtained with 0.2 M (2222 and 0.4 M Na¢4B in PYR. (a) 135.400; (b) 116.4OC . . . . . . . . . . . . . . . . . . . 108 Computer fit of spectra obtained with 0.2 M C222 and 0.4 M Na¢4B in THF. (a) 49.40C; (b) 44.10C. . . . . . . . . . . . . . . . . . . . 109 Computer fit of spectra obtained with 0.2 M c222 and 0.4 M Na¢4B in THF. (a) 56.80C; (b) 62.80C. . . . . . . . . . . . . . . . . . . . 110 Computer analysis of the 23Na lineshape for a solution containing 0.15 M C222 and 0.6 M NaBr in EDA at 25.50C. (a) No first—order phase correction; (b) first—order phase corrected . . . . . . . . . 115 xiv Figure Page + from 33 Arrhenius plot of k (rate of release of Na C222) for EDAl solution. (D) No first-order phase correction; (O) first-order phase corrected. . . . . . . . . . . . . . . . . . . . . 117 34 Plot of attenuated intensity I/IO XE frequency for a 5000 Hz 4-Pole Butterworth active filter . . . . 118 35 Plot of the ratios TObS/Tcalc vs temperature (0C) for EDAl, EDA2 and EDA3 solutions. . . . . . . . . 124 36 Arrhenius plot of k (rate of release of Na+ from C222) for EDA2 solution. . . . . . . . . . . . . . 125 37 Arrhenius plot of k (rate of release of Na+ from C222) for EDA3 solution. . . . . . . . . . . . . . 126 38 Arrhenius plots of k (rate of release of Na+ from C222) for H20, THF and PYR solutions . . . . . . . 127 39 Three possible models for a species of stoichiom- etry M” (other than an alkali anion). All of these models permit solvent interaction with the outer p electrons of the cation. Ammonia is used to represent any amine or ether solvent. . . . . . 132 40 23Na NMR spectrum of a solution of Na+C222 . Na‘ in EA (20.2 M) at 1.4OC. Reference is saturated aqueous NaCl; positive shifts are diamagnetic. . . 135 41 23Na NMR spectra of Na+C222 . Na" solutions (20.1 M) in three solvents. All chemical shifts are refer- enced to aqueous Na+ at infinite dilution. . . . . 139 XV I. INTRODUCTION Sodium and potassium were first observed to be soluble 1 in liquid ammonia by Weyl in 1863. Since that time, the Study of physical and chemical properties of metal ammonia (M—NH3) solutions has been given considerable attention. Properties of these solutions range from electrolytic at infinite dilution to metallic at high concentrations. A number of models have 2-10 been postulated to describe the species that exist in solution at low and moderate concentrations (< 0.1 M). Most everyone agrees that the properties in very dilute solutions (< lO—3M) can be predicted by a two-species model. These two species are the solvated cation and the solvated electron (esolV ). The situation becomes much more complex at con- centrations above 0.005 M, since cation-electron and electron- electron interactions become increasingly important. Some 11,12 properties such as conductance and magnetic susceptibil— 13-15 are concentration dependent and imply that new species ity are formed, while other properties such as the optical spectrum 16—18 ESR absorption line shape,19—21 9,22,24 and extinction coefficient, and partial molar volume of the solute change so little with concentration that formation of distinctly different new species seems to be ruled out. To this date no one model ap— pears to describe satisfactorily the behavior of properties 24 25 of M-NH solutions. Dye has pointed out that a "weak— 3 interaction" model qualitatively agrees with the experimental results, but that the quantitative agreement is poor. 1 In contrast to metal ammonia solutions which show only a single, metal independent ESR and optical absorption band (both attributed to the solvated electron), the ESR and optical properties of metal—amine and metal—ether solutions are much richer in information about distinguishable species. Some of the most obvious differences between metal-amine or metal— ether and M—NH3 solutions are the drastic reduction in solu- 25-28 bility of the former, the identification by ESR of a monomeric species with stoichiometry M and the appearance of 29’30 attributed to a dia- metal-dependent absorption bands magnetic species with stoichiometry M-. This availability of specific spectroscopic information makes the study of these solutions particularly attractive. In spite of these advan— tages, past studies had been severely hampered because of poor solubility of the metals (in many cases the metal does not even dissolve). In 1969 a major breakthrough came from the work 31 of Dye, Nicely, and DeBacker They demonstrated that di- cyclohexyl 18-crown—6 (a member of a class of monocyclic polyether complexing agents first synethsized by Pedersen,32 Figure l) considerably enhances alkali metal solubility in 30,33 ethers. Further investigations disclosed that the complex- ing agent 2,2,2 cryptand (C a member of a class of macro- 222' heterobicyclic complexing agents first synthesized by Lehn 34 Figure l) enhances metal solubilities in amines 2221-. and ethers even more effectively than dicyclohexyl 18—crown- 6. The use of these synthetic complexing agents extends the solvent range in addition to increasing the concentration of 2,2,2 C rypta nd (C222) (“AI (3 C) C. .3 C3.) 18-C row n— 6 08-0-6) Figure 1. 2,2,2 Cryptand and l8—crown—6. dissolved metals in solvents which dissolve metal unassisted. The research efforts described in this thesis are focused on the nature of the species responsible for the metal— dependent visible absorption band in metal-amine and metal— ether solutions. The evidence to date indicates that this species is diamagnetic and has the stoichiometry M-. On the basis of optical evidence alone one cannot distinguish among several species with stoichiometry M“. These are: (1) an ion-triple between a cation and two spin-paired electrons, e— ‘ M+ ° e_, (2) an ion-pair between a cation and a di— electron, M+ ° eg-, (3) a cation with a pair of electrons in an expanded orbital which includes the first solvation sheath molecules, and (4) a genuine alkali anion. A method which can distinguish among the various models is alkali metal NMR. A genuine anion M_ with two electrons in an outer s orbital should have considerably different NMR properties than the corresponding cation M+. Chapter VI includes a summary of a multinuclear alkali metal NMR study of metal—ether and metal- amine solutions which contain Na—, Rb— and Cs_. The results strongly favor the alkali anion model. A detailed descrip— tion of the first synthesis of a crystalline salt of the sodium anion is included in the experimental section (Chapter IV) . This crystalline salt of the sodium anion and the alkali metal NMR results are the strongest evidence that genuine alkali metal anions do exist.35_38 The key factor which makes alkali metal NMR studies of metal solutions feasible is the enhancement of metal solubility which is provided by cryptands and crowns. Prior to 1970 the maximum concentrations of sodium in amines and ethers were less than 0.01M. These concentrations were too low for prac- tical NMR studies. With the aid of C solutions as con~ 222' centrated as 0.4M in sodium metal can be prepared. A second goal of this research is focused on the applica— tion of pulsed 23 Na Fourier transform NMR lineshape analysis to permit quantitative examination of exchange rates of sodium cations in the presence of C222. Initially it was felt impor- tant to examine the effect of sodium cation exchange upon the observed lineshape of a sample which contained a dissolved sodium salt and half the stoichiometric amount of C222. The overall chemical equilibrium is given by reaction 1. + * + k * + + Na C + Na 2 Na C + Na 1.1 222 222 Sodium cations undergo an exchange between bound and solvated sites which can affect the NMR lineshape. The influence of exchange was first investigated in order to determine whether 23Na NMR studies of metal-amine and metal-ether solutions would be feasible. Exchange rates39 have been reported from PMR studies on D20 solutions which contain Na+C222 cryptate. It was clear from the results of this study sodium cation exchange may be slow enough to be observed directly by using 23Na NMR. A preliminary investigation by Ceraso and Dye4O did indeed show that exchange rates are slow in ethylene- diamine (EDA) at 25°C. This is the first example of sodium cation exchange which is slow enough to exhibit clearly de— fined separate signals for two environments. Since the study of such cation exchange phenomena was relatively new, we decided to examine the influence of the solvent on the ex- change rates. Chapter V describes an in—depth study of the . . . . + . kinetics of complexation reactions of Na —C complexes in 222 several solvents. This chapter also includes a discussion of some of the causes of systematic errors which are common to pulse Fourier transform NMR techniques. 23Na NMR lineshape analysis was done in Since some of the collaboration with Patrick B. Smith, his M. S. thesis should also be consulted for further details. II HISTORICAL 2.1 Metal Ammonia Solutions Alkali and alkaline earth metals dissolve in liquid am— monia to form blue-colored, paramagnetic solutions. A wide range of concentrations exists due to the high solu— bility of metals in ammonia. In the very dilute concentra- tion range the solutions behave as binary electrolytes with solvated cations and solvated electrons as the ionic species. Complex association of these species occurs in the inter- -l mediate concentration region (10"3 — 10 M). These inter- actions influence the physical and chemical properties and complicate our understanding of the solutions. Above 1 M, a non-metal to metal transition occurs. Solution properties become metallic and the conductivity of saturated solutions approaches that of the free metal. In the description of the properties and existing models of M—NH solutions, attention will be mainly given to solu- 3 tions with dilute and intermediate concentrations of metal. It is not the author's intention to give a complete history of the prOperties of these solutions. A comprehensive history is contained in the Proceedings of Colloque Weyl 41 42 43 44 I I I I, II III and IV. The author merely intends to focus on selected and current tOpics of interest. 2.1.1. Species in M-NH3 Solutions A variety of models have been postulatedz-10 to des- cribe the species at low and intermediate concentrations. Species with stoichiometry M+, e—, M, M—, eg— and M2 have been postulated to exist within these models. To add to the complexity of these models, the nature of species with the same stoichiometry may vary according to the model. The solvated electron as a species is well established and will not be directly considered in this discussion. 2.1.1.1. Evidence for Monomer Species in M—NH3 Solu- tions - Certain properties of metal solutions indicate the presence of cation—electron interactions. The strongest evidence comes from the electrical conductance of M-NH3 11’12 and transference measurements45 solutions. Conductance yield the concentration dependence of the cation conduc- tance. The cationic equivalent conductance decreases with increasing concentration of metals, similar to ordinary electrolytes in ammonia. The total equivalent conductance also decreases with increasing concentration of metal and shows a pronounced minimum at 0.04 M concentration of metal. A single equilibrium constant, K suffices to describe the 1’ concentration dependence of the conductance below about 0.005 M. Furthermore, the magnitude of K1 is similar to that observed for normal 1 to l electrolytes which form ion—pairs in ammonia. 46-48 Cation NMR spectra also show the effects of cation-electron interactions. Concentration dependent paramagnetic (Knight) shifts are observed. The Knight shift of a nucleus is caused by the hyperfine interaction between unpaired electrons and the nucleus. This interaction also contributes to the spin-lattice relaxation time of the metal 11 sec.)48 nucleus. The correlation times are very short (10— indicating a short lived cation—electron interaction. A single extremely narrow ESR line (assigned to the solvated electron, e— solv.) 18 observed in dilute solu- 19-21 tions. The band is structureless and has a g-value of 2.0012. No hyperfine interaction with the metal is ob— served. Metal solutions also exhibit a broad optical absorption band (assigned to e— solv ), which peaks at 0.85 e.v. and is independent of solute. The band shape and molar extinction coefficient are independent of concentra- tion and the position of the peak maximum is only slightly 16-18 Cation—electron interactions concentration dependent. strongly influence the conductance and NMR properties whereas they have almost no effect upon ESR and optical properties. Recent pulse radiolysis studies of liquid deuterated ammonia solutions which contain dissolved sodium and potas— sium amides show at least two optical absorption bands.49'50 The initial transient (Amax = 1500 nm) is extremely short— lived (<150 usec.) and is assigned to the solvated electron at infinite dilution. The residual absorption (Amax = 1640 nm.) is shifted to the red and is interpreted as represent— ing two or more overlapping bands corresponding to esolv 10 and a metal—electron species. However, from pulse radiolysis experiments performed on mixed ammonia—amine solutions, one would expect that ion-pairing effects would shift the - 50 peak max1mum of esolv. towards the blue. 2.1.1.2. Evidence for Diamagnetic Species in M-NH3 Solutions - Magnetic data offer the most convincing evi— dence that at least one diamagnetic species is formed in M—NH3 solutions. Static magnetic susceptibility studies 14 performed by Freed and Sugarman and radio-frequency ESR 15 both show a studies performed by Hutchison and Pastor dramatic decrease in molar paramagnetic susceptibility with increasing metal concentration. At 0.1 M metal concentra- tion, the electrons are over 90% spin—paired. The molar paramagnetic susceptibility is strongly temperature de— pendent with the diamagnetic states preferred at lower temperatures. Susceptibility data for potassium indicate a temperature dependent spin—pairing enthalpy, whereas 15 51 data for sodium do not. Studies by DeMortier et a1. indicate that electrons in sodium—NH3 solutions are much less spin—paired than those in K-NH3 solutions. However, 52 53 indicate that spin-pairing studies of Lok and Tehan is the same for both K-NH3 and Na—NH3 solutions. More experimental work is needed to determine whether or not the cation plays an important role in the spin—pairing reaction. An extremely narrow ESR line is observed in M-NH3 11 solutions up to concentrations where the electrons are 90% spin—paired. The narrowness of this line requires that reactions of the type 2 D + S in which D can be any doublet state (such as e— or M) and 2- 2 I times longer than a microsecond. Since magnetic inter— 8 any singlet state (such as e M- or M2), have life- actions resulting from collision complexes or long-range interactions would not be expected to exist for more than a few nanoseconds, it seems appropriate to refer to the presence of a diamagnetic species. Enthalpy measurements for Na-NH 22 3 solutions performed by Gunn parallel the change in magnetic susceptibility Of K-NH3 solutions with both concentration and temperature. The enthalpy data can be described satisfactorily by the following reactions: where K2 comes from potassium magnetic susceptibility data and K1 is determined from sodium-ammonia conductance data.24 The slight red shift in the optical absorption band with increasing concentration of metal is the only other property 12 which correlates with the concentration dependence of the magnetic susceptibility.16.l8 Several diamagnetic species with different stoichiometries have been postulated to exist in these solutions. The nature of this diamagnetic species will be considered in a later section. 2.1.2. Models for Metal Ammonia Solutions 2-10 to describe A variety of models have been postulated the species that exist in M-NH3 solutions. The strength of any of these models depends on how well they can predict all of the known properties of these solutions. Dye24 has clas— sified the models under one or a combination of three gen- eral categories: (1) metal-based species; (2) double oc- cupancy of cavities; (3) electrostatic aggregates. Models classified under metal-based species include those .2’4’6’7’10 The for— which form species such as M, M_ and M2 mation of these species results from strong interactions between cation and electrons (stronger than simple electro- static interactions between charged ions). Double occupancy models3’54’55 consider a pair of electrons, e§_, in the same polarization center (or cavity). A number of calculation556_62 have been done to determine whether or not two solvated electrons at separate sites would be more or less stable than two spin—paired electrons trapped at the same site, but the results tend to oscillate between the stability of one over the other. In any event, consideration of the ion- pairing interaction for a normal 1—2 electrolyte in ammonia 13 shows that the fraction of free e3- relative to the ion— + 2‘ (stoichiometry M”) will be negligibly small pair, M - e2 even at concentrations where spin—unpairing is complete. Both metal-based models and double-occupancy models have a common weakness. They both predict the formation of new species even though there is no observable change in the optical spectrum and molar volume of the solute. The final category considers the formation of ionic aggregates between M+ and egolv without destroying the basic 8,9 characteristics of the species. Species with stoichiometry M, M- and M are viewed as ionic clusters formed from the 2 fundamental species M+ and e— solv.‘ The PrlnC1p1e advantage of the ion cluster model is that the optical spectrum and molar volume of the solute are preserved. However, electro— static considerations alone, do not permit concentrations of triple ions and ion quadrapoles in high enough concentration to explain the magnetic susceptibility or enthalpy data. While not all models claim the same stoichiometry, M+, e', M, M— and M2 include simpler models as special cases. Before arguing the merits of one structure over another with species of the same stoichiometry, let us note that Dye24 has tested the stoichiometry represented by these species to determine whether it agrees with the experi- mental data. The choice of equations used to describe the equilibria among all the species for any model is somewhat arbitrary. The following set is complete and independent: 14 K _1 (1) M++e :14 -K2_ (2) M + e Z M K _ 3 (3) M +M+ZM2 Values for Kl were calculated from Na conductance data and values for K2 were calculated from potassium magnetic sus— ceptibility data. Because both are ion—pairing reactions, K3 was assumed to be equal to K1' With these values of K1' K2 and K3, activity coefficients and transference numbers were calculated and compared with measured values. The re- sults indicate that reaction 1 alone satisfactorily describes these electrochemical properties. Inclusion of reactions 2 and 3 gave large deviations from measured behavior.24 In conclusion, none of the present models for concentra— tion dependence of the properties of M—NH3 solutions are able to describe completely the experimental results. Dye has suggested that a "weak-interaction" model involving normal electrostatic interactions between two species M+ and e‘ is attractive.24 The concentration independence of partial molar volume of M and the optical spectrum of egolv. will be preserved with this type of model. In ad— dition, the electrochemical properties will also be described by normal electrostatic interactions which include ion- pairing interactions. As for spin-pairing, a long range electron—electron interaction is assumed and this should have little or no effect upon the electrochemical properties 15 or the optical spectrum. To explain the long lifetime of the diamagnetic species, the following type of exchange process can be invoked: e— + M- + M— + e— In this case one electron in the singlet state is replaced by another. This need not lead to ESR line-broadening. Although qualitatively correct, this model lacks quantita— tive treatment. 2.2 Alkali Metals in Amines and Ethers 2.2.1. General Properties Alkali metals dissolve in amines and ethers to a much lesser extent than in ammonia (in many cases the metals do not dissolve). In spite of the low solubility of metals in amines and ethers, studies of these solutions are rich in information about distinguishable species. The optical spectrum consists of two bands, one in the visible and the other is near the infrared (IR). The posi— tion and width of the IR band is solvent dependent and nearly metal independent. In contrast, the visible band is only slightly influenced by solvent and is strongly metal dependent. The species responsible for the metal independent IR band has been identified as the solvated electron. The identification is based upon a comparison 16 of IR bands in metal solutions with those produced by pulse 63-68 The species responsible radiolysis in pure solvents. for the metal dependent visible absorption band was not clearly identified until after 1968. Prior to this date, there was much discrepancy about the nature of the species responsible for the visible band. Experimental results in one laboratory could not be reproduced in another. The chaos 69 demonstrated that ended after Hurley, Tuttle and Golden potassium ions in solution could exchange with sodium in borosilicate glass. It quickly became clear that for a given metal, at most two optical bands need be considered and that the visible band could be associated with a species of stoichiometry M—. The ESR spectrum shows a narrow singlet (assigned to the solvated electron) with an additional hyperfine pattern superimposed upon it.25'26 The hyperfine structure comes from a strong interaction of the electron spin with the nuclear spin of the metal. This observation gave the first unambiguous identification for a species with stoichiometry M (referred to as a monomer or ion pair). 2.2.2. Major Species Involved in Metal-Amine and Metal- Ether Solutions 2.2.2.1. Solvated Electrons — The strongest evidence that solvated electrons exist in metal—amine and metal— ether solutions comes from the comparison of the IR band with the spectrum produced by pulse radiolysis of the pure 17 solvent. The band shape and peak position for both ammonia and ethylenediamine are independent of the method used. With the discovery that "crown"31 and "cryptand"30'33 can be used to dissolve metals in a number of solvents and that an excess of cryptand yielded the IR band, it was possible to compare the IR band in metal solutions with those produced by pulse radiolysis and flash photolysis. Extensive comparison of 1.30 _——— this was made by Lok et_ and the agreement between the two methods was good. However, the agreement is not exact 70 that the metal solution bands are, and it has been noted in general, broader and at slightly higher energies than those found by pulse radiolysis. The variation in width and peak position is probably due to the higher concentrations of metal solutions and may be attributed to formation of a second diamagnetic species (other than M_) or to an ion— pairing effect. The single narrow ESR line with the same g-value as that of M—NH solutions also supports the existence 3 in these solutions. Additional evidence comes from 71-73 Of eEolv. the conductance of cesium—ethylenediamine (EDA) solutions. When Walden's rule is used to account for viscosity differ- ences between ammonia and ethylenediamine, the limiting con- ductance of Cs—EDA and Cs—NH3 solutions is nearly the same. Cs—EDA solutions do not show any visible absorption band and, hence, the conducting anions must be the same species which are responsible for the IR band. 18 2.2.2.2. Monomers - The first evidence of the existence of a monomeric species, M, was obtained by Vos and Dye25 from ESR studies of rubidium and cesium methylamine solutions and independently by Bar-Eli and Tuttle74 from ESR studies of potassium in ethylamine solutions. A marked increase in hyperfine Splitting with a decrease in solvent polarity and 75 The intensity of an increase in temperature is observed. the hyperfine signal is low and indicates that the monomer species is only a minor constituent of the metal-solution (no Optical spectrum is observed for the monomer species in metal solution under equilibrium conditions). Other evidence for the existence of monomers comes from flash photolysis and pulse radiolysis of metal solutions (in these cases an optical spectrum for the monomer is detected, but only under transient conditions). A number of flash-photolysis studies have been performed on metal-ether 74—81 Most of the solutions and metal-amine solutions. studied contain M+ and M— as major species. When a solution is flashed with light at a wavelength corresponding to Amax M— (band maximum of M-) photo bleaching of this band occurs with subsequent formation of a broad IR band. This IR band transforms rapidly to an intermediate absorption in the visible and then the intermediate band decays slowly back to the original M“ band. Even though earlier studies had been done by others, Kloosterboer et 31.81 were the first to demonstrate that the intermediate absorption band has the 19 stoichiometry M. These results are consistent with the following reaction sequence hv M+ M + M+ + 2e“ + 2M+ ' ’ ' e + M + M+. Pulse radiolysis studies also indicate the existence of monomers (again under transient conditions). Bockrath and 65 reported the spectrum of the sodium-electron ion 82 Dorfman pair (Na) in THF. Pulse radiolysis studies of ethylamine- THF mixtures show only a single monomer band and the peak position lies at wavelengths intermediate to that observed in the pure components. A correlation is observed between the shift in Amax of the monomer and the magnitude of the hyperfine splitting constant in these solutions.82 As the percent atomic character increases (increased hyperfine splitting constant, corresponding to a larger contact density at the nucleus) the Optical spectrum of the monomer shifts increasingly toward the blue. 2.2.2.3. Spin-paired Species — The existence of a spin- paired species, other than M", in these solutions has not been shown. Studies of solutions of K in ethylamine-ammonia mixtures with spin concentrations of the order of 10"6 M indicate a total concentration for the IR absorbing species to be about 10‘4 M.70 If these results are reliable, the IR absorption must include contributions from both paramagnetic 20 and diamagnetic species other than K“, since the band maximum of K- would appear in the visible. If other diamagnetic species besides MI exist in these solutions, then it may be possible to correlate the magnetic properties of this new species with the spin—paired species that exists in M-NH3 solutions. 2.2.2.4. Alkali Metal Anions — Chemists may view the existence of alkali metal anions with skepticism (this was especially true at the beginning of this research, since no known salts of alkali metal anions existed), but the evidence for the existence of these species has mushroomed in the past six years. The assignment of the stoichiometry M“ to the species responsible for the metal dependent visible band is based on numerous experimental facts. That alkali anions exist in the gas phase was shown by 83 Metal solutions which Russian scientists in the 1960's. yield a large visible band and no IR band have very weak or no ESR signals. The optical spectra of these solutions, unlike M-NH3 solutions, exhibit metal—dependent absorption bands. Figure 2 shows the spectra of egolvo' Cs“, K‘ and Na— in tetrahydrofuran at 250C in the presence of either crown or cryptand. It seems unlikely that normal electrostatic interactions between cation and electrons can cause such a large shift and metal dependence of the solvated electron absorption band. By comparing the temperature and solvent 21 .csouo no pcmummuo mo mocmmoum may SH nee EH Ass .42 use Ame Is .lmv Imo .AHV >Homm mo 60mm pm muuomem .m masons aflIO—X—IEUV genus—3:05:92, ca 2 2 .1 m. o. m e e A _ _ a _ _ I4 _ . I.~an Iiexnv v. m” NV p uv .IAVAVmw W . D .Imwnvya 22 dependence of the optical bands in metal-amine solutions with the charge—transfer-to-solvent bands of I—,84_86 Matalon, Golden and Ottolenghi29 assigned the metal-dependent band to the alkali anion. Some of the characteristics of CTTS trans— itions are: 1. Pronounced dependence of Ama upon solvent. X 2. Shift of Ama to lower energies with a decrease in x temperature. 3. Correlation between the shift of Amax with solvent and the temperature coefficient of Amax' 4. Correlation between the position of Amax and the size of the anion, with a shift to lower energies for larger ions. All of these characteristics have been observed for M— bands. Lok, Tehan and Dye30 have made an in-depth study and have shown excellent correlations between the peak positions of Na' and K" as shown in Figure 3. CTTS theory, therefore, strongly suggests that the species responsible for the visible band is an alkali anion, but it does not prove this. For example, solvated electrons also have many of the same CTTS characteristics as the M‘ species. This is shown in Figure 4 by the excellent correlation of the solvated electron peak position with the K‘ peak position in various solvents. Thus, it might be expected that a species like e_ ' Na+ ' e- (triple—ion) should give rise to spectra with CTTS characteristics. Additional evidence Na'PEAK posmon Figure 23 EDA6// I2PDAo EA'O ODIGLYME °DME OTHF HMPA DEE ONT 0 I l 3. 1.1 1.2 ( C M_lx |0_3 ) K" PEAK POSITION The relation between the peak position of Na- and of K" in diisopropyl ether (DIPE), diethyl ether (DE), hexamethyl phosphoric triamide (HMPA), tetrahydrofuran (THF), dimethoxyethane (DME), diglyme, ethylamine (EA), 1,2-propanediamine (PDA) 1.3 and ethylenediamine (EDA) at 250C. Ea >H0m .oOmN um mucm>H0m mdoflum> Io mo.ocm Ix mo coeuemom some on» cmmzhmn coeumHmH one .e ousmflm 20:50.. 52: L. r. 9x753 See. . S E S 24 E _ _ 1 a l v )XOWA [— OlXVVD c... I -o ( noulsoa )IVJd "“9 wH wmnocm on mcflocommmuuoo Ednpoomm Hoveuosuommm ADV .Hoouo umnfim cfl moflmuoco cmEooN Hmoaosc com: mcflamsoo DHOQSHUMSU mo uommwm Amy .m oudmflm Q m IILIIINE 0T; ‘— 31 NaNO391 (hexagonal crystal structure). The NMR spectra of crystals such as the alkali halides do not exhibit quadru- pole splitting since the crystals have cubic symmetry and no field gradient exists. In solution, the effect of quad— rupole splitting is not observed since for normal liquids, the time scale of solvent fluctuations (which produce field gradients) is 4 to 5 orders of magnitude shorter than the time scale corresponding to the quadrupole coupling. In this chapter we will describe some of the general features of alkali metal ion NMR. The description is by no means complete. It is only meant to give the necessary background for the NMR studies described in Chapters V and VI. 3.2 Shielding Constants of Alkali Metal Ions The shielding constant, 0, is defined by mo = y(l—0)HO, where Y is the gyromagnetic ratio of the nucleus and H0 is the external magnetic field. When the gaseous atom is taken as the reference state, then 0 = 0 and y refers to that for the gaseous atom. A negative value for o is inter- preted as a paramagnetic (low field) shift relative to the reference state and correspondingly a positive value for o is interpreted as a diamagnetic (high field) shift rela- tive to the reference state. Note that this sign is op- posite that frequently used to express chemical shifts but its use provides an internally consistent sign conven- tion. 32 3.2.1. Alkali Metal Ions in Crystals The shielding constant of alkali metal ions in crystals is paramagnetically shifted (shifted to lower field) from that of the free gaseous ion. In crystalline salts the chemical shift also depends upon the counter anion.92 The factors governing these shifts are qualitatively under— stood, but are not well accounted for quantitatively. During the late 1950's and up to the mid 1960's a number of attempts had been made to calculate paramagnetic shield— ing constants for alkali metal ions in solid alkali halides. The starting point for all these calculations was Ramsey's expression93 for the total shielding of a nucleus, given by 2 r i—f f k k k 0 = ( e )[(9 I) -——-————-|w ) 2mc2 O k rk3 O + 2 (E0 - Em)"l{(w | X Ek | Wm) m k E A k k (wm I Z ;_3 lqlo).+ (Yo I E ;_3 I Wm) k k (9m I E 1k I NO)}], 3.1 where To and Wm represent the many-electron ground- and excited—state wave functions, EO and Em are the ground- and excited-state energies, Ek is the angular momentum 33 th electron and rk is the radial distance of the kth electron from the origin at the nucleus. operator of the k The first term in Equation 1 is the diamagnetic con— tribution to the total shielding. For a gaseous alkali metal ion, the diamagnetic shielding, is proportional 0D, to the magnetic field produced at the nucleus by induced currents which arise from the Larmor precession of electrons in the external magnetic field. The second term is the paramagnetic contribution to the total shielding and is zero for an isolated gaseous alkali metal atom or ion. A paramagnetic shift is produced by introducing orbital an- gular momentum into the wavefunction. For alkali metal cations in alkali halide crystals and in solution, the paramagnetic contribution to total shielding is not zero. In the crystal, the wave functions are distorted from those of the hypothetical ideal crystal (considered as combina— tions of distinct positive and negative ions, each con— stituent ion having a spherically symmetric closed shell electronic configuration identical to that expected for the isolated ion, with the complete assembly being held together by electrostatic forces) and hence orbital angular momentum is introduced and produces a paramagnetic shift. As can be seen from Equation 3.1, both 0D and Up are tensor quantities in the general case. For all the systems discussed, cubic symmetry is assumed and the scalar part will be used. In order to evaluate op, a knowledge of Wm is needed. Since in almost all cases, W is not known, In 34 Ramsey93 introduced the average—energy approximation. With the assumption of cubic symmetry and the average energy approximation, the total shielding is now given by a= (ill—III) 3mc2 O rk 2 E, + e22” I ”EkTI Io“ 3'2 Am c kk' rk where A is the average excitation energy. Two models have been proposed for the origin of mag— netic shielding based on the perturbation—theory approach of Ramsey. The first is the charge~transfer—covalency model 94 due to Yosida and Moriya, and the second the overlapping- ion model by Kondo and Yamashita (KY).95 Ikenberry and Das96 have pointed out that calculations of quadrupole coupling constants in diatomic—alkali-halide molecules clearly in- dicate that there is little, if any, covalent bonding of the charge-transfer type in these crystals. Thus, in the most recent calculations, only the KY model has been con— sidered. The KY model assumes that short range repulsive forces between adjacent ions cause a change in the op value of M+ upon going from the gas phase (op = 0) to the nonideal crystalline state. From the KY model only overlap between outermost s and p orbitals for both positive and negative ions is considered. Since inner—core orbitals are tightly bound and do not extend appreciably into the region between 35 the ions, they remain spherical and are not considered. Also calculations show that the change in CD for M+ gas + M+ crystal is two to three orders of magnitude less than 96 + + op. Thus, for crystal 0 (Mcrystal YE Mgas) ~ op (assum_ ing a reference state 0 = Up a 0 for the gaseous metal ion). Under these conditions with the application of Equation 3.2, Ikenberry and Das96 have calculated chemical shifts for Rb+ ions in halide crystals. Their final expression for O (M+ 97 + - u p crystal is. Mgas) is given as follows 160I2 _ < —3> [ISOSIZ + ISOOIZ + ISUWI Op ' A ri p ij ij ij CO TT'TT where d is the fine structure constant, p is the ex— pectation value for an outer p—electron of the central ion i, Sij are the two—center overlap integrals between outer p orbitals of the central ion i and the outermost s and p orbitals of other ions j, s represents an s orbital, and o and n are p orbitals parallel with the line of nuclear centers and perpendicular to the line of nuclear centers respectively. In summary, the features that are important for mag- netic shielding as given by Equation 3.3 are: 1. The expectation value p for the outer p orbital of the central ion. 36 2. The average excitation energy A, usually taken to be the energy of the np + np + 1 electronic transi- tion of the central ion. 3. Overlap between neighboring ions, which depends upon the internuclear separation. The average excitation energy determines the amount of ex— cited state p character introduced into the wavefunction. This causes a deformation from spherical symmetry and produces the paramagnetic shift. Some values of p and A along with their ratios are shown in Table I. Ex— perimentally, the chemical shift range increases as atomic number increases and this increase is in agreement with the increasing values of % (ri3>p' At the time Ikenberry and Das, and Hafmeister and Fly- gare performed calculations for magnetic shielding of ions in crystals, there were no experimental values available for 0(M+ vs M+ ). It has been only within the last crystal —— gas few years that these values could be obtained. Experimental values were determined from a combination of precision 99-101 102 103,104 I NMR atomic beam, and optical pumping experiments. From these measurements it is now possible to obtain values for o(M+ vs M ) and since many determina— H20 —— gas + + I f - — tions 0 0(Mcrystal vs MHZO) have been reported, an experi + . mental value for 0(Mcrystal XE Mgas) can be obtained. How ever, calculated values of 0(M+ crystal) from theory are referenced to the gaseous ion. The difference between 37 Table I. Values of the average excitation energy A, and the expectation value p for alkali metals.a -3 A 1 p l —3 Ion a.u. Rydbergs A p Na+ 16 2.72 5.9 K+ 12.94 1.62 7.98 Rb+ 20.22 1.47 13.8 Cs+ 23.42 1.25 18.7 a) All values taken from Reference 98. 38 + crystal 35 + Mgas) and 0(M vs M + + crystal —— ga (M 0(M D gas S) is o 23 Mgas) which is the diamagnetic contribution caused by the addition of an electron to a gaseous ion. The term GD (Mgas 23 M ) can be reliably calculated from accurate gas atomic wavefunctions by applying the first term in Equation 3.2. Values for these diamagnetic shielding constants for Li, Na and K are less than 5 ppm and tend to get smaller 105 Thus, for Rb and Cs, the . + + + . assumption 0(Mcrystal gs Mgas) ~ 0(Mcrystal gs Mgas) is as atomic number increases. reasonable. Comparison of the chemical shift values cal- 106 culated by Hafmeister and Flygare, and Ikenberry and Das96,97 + With the measured values of 0(Mcrystal XE Mgas) are given in Table II. The agreement is qualitative only. However, it should be pointed out that theoretical values for the change in shielding with pressure for alkali halide 96:97,106 107 are in much better crystals and for xenon gas agreement with the experimentally observed chemical shift changes. From these calculations it appears that the KY model offers a reasonable explanation of magnetic shield— ing. 3.2.2. Alkali Metal Ions in Solution Solvated cations also undergo large paramagnetic shifts from the gaseous ions as shown in Table II for rubidium and cesium. Changes in the diamagnetic shielding as a function of solvent are expected to be small compared to changes in op.109'llo Thus, as for the crystals, changes 39 .NOH moawnwmmm he .woa mocmummom A0 .mm mucoummom An .NOH 8:8 mos .Na mmocmnwwwm eons Ammo: ms em fineszvmxmo new .iem HHUIHz m> we powzvmxoo .Awm new: m> mhnwzvmxoo mcflfifidm an omcflmuno DMD: modao> Hmpcmfifluwmxm Am comm- ems- Hmo +mo comm- mam- ammo +mo comm- amm- Home +mo em.eem- fine 8 on comm- ewe- emu +mo coma- mam- Hem +nm Dome- mmm- umnm +nm UOMHI mHmI HUQM +Qm ooeHI cam- mam +nm ee.HHNI fine 8 on seam- mmm- umnm +nm Ammo: .m> .wwwo ucm>aom mmmmmvm mmMvam wofluqu coH >Homzvmxoo A +2 .m> MA 2 .m> + mwuwzvoamoo mhuwzvmxmo .coHu9H0m msoosom one mampmwuo ocflams EH mQOH Edemoo one Eseoensu mom mamanm Hmoflsmco oopmasono tom Hmucoswuomxm .HH Danae 40 in magnetic shielding are reflected predominantly through changes in Up. An interesting correlation between the Gutmann donor number (D.N.) and the relative chemical shift of sodium 111,112 ions in different solvents has been reported. The Gutmann donor number of a solvent is a measure of its Lewis 113 As the D.N. increases, the solvent donor basicity. ability increases and an increasing paramagnetic shift of the sodium cation is observed. A linear correlation be— tween Gutmann's donor number and the relative magnetic shielding of sodium ion in different solvents is observed. For sodium, the solvent effect on the magnetic shield— ing occurs primarily in the first solvation or coordination layer of the sodium cation. If a sodium is trapped within the cavity of a complexing agent such as C222 (the first coordination layer of solvent is completely replaced by C222), the chemical shift of the cation is nearly solvent independent. On the other hand, the free sodium ion chem— ical shifts change appreciably in these solvents as shown by Table III. The same behavior is observed for lithium ion complexed by a smaller cryptand ligand.114 Calculation of shielding constants for alkali metal ions in solution is a much more difficult task than for ions in crystals. As a consequence, only one semiquantita- tive calculation based upon the KY model has been performed for Rb+ in aqueous solution. The theoretical value for 0(Rb; XE Rb;a ) = -65 ppm is in poor agreement with the 20 s 41 Table III. Chemical shifts of free and cryptated sodium ions in solution. + + Concentration 0(Nasolv VS Nam-dil—HZO) (M) Solvent (ppm) 0.2 NaI H20 -0.1 0.6 NaBr EDA —13.7 0.4 Na¢4B THFa +7.4 0.2 Na¢4B PYRb -0.9 + .— 0.2 Na C222I H20 +8.5 0.3 Naiczzzsr‘ EDA +10.6 4. .. 0.2 Na C222¢4B THF +11.9 0.2 Na+C222¢4B— PYR +12.4 a) THF = tetrahydrofuran. b) PYR = Pyridine. 42 experimentally determined value 0(RbH20 XE Rbgas) = -212 ppm.97 However, the model used for the calculation is one in which a Rb+ is surrounded by six water molecules with the oxygen end oriented towards Rb+ and the O-H bonds point— ing away. The wavefunctions for H20 were based upon a model of localized electron-pair bonds. In view of the uncertainty of the model and the crudeness of the wavefunctions, the poor agreement between theoretical and observed values is understandable. It seems that Equation 3.3 at least quali- tatively predicts the effects of a paramagnetic shift for ions in solution. 3.3 Quadrupole Relaxation of Alkali Metal Ions in Solution The magnetic relaxation of nuclei with spin greater than 1/2 is almost always caused by the interaction of their electric quadrupole moment with the electrical field gradient at the nuclear site.~ Thermal motion of ions and solvent molecules can produce fluctuating electric field gradients at the nuclear site. The time evolution of these fluctuations contain Fourier frequency components which are at the Larmor frequency of the alkali metal ion and hence can induce a transition through coupling with the quadrupole moment. The quadrupole relaxation rate for a nucleus in the motionally narrowed limit (wTC< noes ou 52 Methylamine and ethylamine were dried over calcium hydride and then vacuum-distilled into a vessel contain- ing Na-K alloy. If a blue solution did not form, the sol- vent was repeatedly vacuum distilled into vessels containing fresh Na—K alloy until a stable blue solution did form. Once a stable blue solution formed (usually no observation of color change after one to two days is a good criterion for stability) the solvents were distilled into Pyrex stor- age bottles. Pyridine was refluxed over granulated barium oxide and then fractionally distilled in a nitrogen atmos- phere. 4.4 Preparation and Purification of 2,2,2 Cryptand During the progress of this research, 2,2,2 cryptand was not commercially available and had to be synthesized by following a brief outline in the literature.34 The synthesis was carried out in collaboration with M. T. Lok and F. J. Tehan. Several major improvements in the pro— cedure were developed by us in the process of synthesizing the material. The details are reported in the Ph.D. thesis of M. T. Lok52 and will not be described here. However, an improvement of the purity of one of the precursors is described. Triglycolic acid (HOCOCHZOCH CHZOCHZOCOH) is a pre- 2 cursor in the synthesis of the 2,2,2 cryptand molecule. The diacid is produced by oxidation of the corresponding 53 glycol with 60% nitric acid and a vanadium catalyst (NH4VO3). Many by—products are formed in the oxidation and it is dif— ficult to separate the acid from these by-products. The following procedure, partly based on a synthesis reported 126 I by Micovic, enables us to obtain triglycolic acid with' reasonable purity. 4.4.1. Synthesis of Diethyl Ester of Triglycolic Acid Add 256 gms of crude diacid to a one neck, 2 liter flask equipped with a magnetic stirring bar, a downward condenser and a collection pot containing 253 gms of K2C03. Add 570 ml of 100% ethyl alcohol, 285 ml of reagent grade toluene and 1.3 ml of reagent grade H2804 to the flask. The flask is heated and an azeotropic mixture of ethanol, toluene and water distill at 75°C. The heating is continued until 300 to 400 ml of distillate are collected and then it is suspended. The distillate is shaken well with K2C03, next filtered through a fritted glass disk, and is poured back into the reaction flask. The mixture is heated again until 95% of the volatile distillate is removed. The solu- tion is allowed to cool slightly and then it is vacuum dis- tilled (without fractionation). Next the liquid is redistilled with an efficient frac- tionation column. The pure diethyl ester boils at 142° under a pressure of 4 mm Hg. The PMR spectrum of the neat liquid consists of a triplet (-CH3) St 1.3 ppm, a singlet u (~OCH2CH20) at 3.8 ppm, a singlet (~C—CHZ—O) at 4.2 ppm 54 and a quartet (-O-CH2-CH3) at 4.25 ppm. 4.4.2. Hydrolysis of the Diester Add 130 ml of diethyl ester, 130 ml of distilled water and 1.3 gms Dowex 50-WX-2 ion exchange resins to a 3-neck 500 m1 flask. The solution is heated to a temperature of approximately 90-95°. The diester is not 100% miscible in water and two phases appear at the start of the reaction. After an hour of heating the solution becomes one phase and liquid starts to boil off between 70-95°C. The distillate is a mixture of water and ethanol. When all the ethanol is used, the boiling point of the distillate goes up to 100° and the hydrolysis is completed. Next, the residual solu— tion is filtered to remove the ion exchange resin and then the remaining water is stripped off with a roto-evaporator. The viscous liquid is placed under vacuum until it solidifies. The PMR of the diacid in D20 shows two sharp singlets with an area ratio of 1:1 and the melting point range is 65—75°C. 4.4.3. Purification of 2,2,2 Cryptand 2,2,2 Cryptand is purified by recrystallization in hexane followed by a vacuum sublimation. If the crude 222 cryptand contained precursor impurities such as the mono- / CHZCHZOCHZCHZOCHZCH2\ CHZCHZOCHZCHZOCH2CH2 recrystallization was found to be effective for separating cyclic diamine, a major portion of this impurity. However, some residual 55 diamine (approximately 5%) could always be detected by PMR studies. The 2,2,2 cryptand in this work gave a PMR spec— trum that did not show impurities and had a melting point range of 70.2-71.1°C. 4.5. Synthesis of a Crystalline Salt of the Sodium Anion The first authentic salt containing a sodium anion was prepared by the following procedure: A Pyrex vessel of the design shown in Figure 9 was used to prepare polycrystalline samples of NaC+ ' Na- in which C is 2,2,2 cryptand (see Figure l). A convenient amount (50 mg to l g) of C was placed in compartment B and sodium metal was placed into the side-arm shown which was then sealed off. The vessel was evacuated to “:10-6 torr and the sodium was distilled into compartment A. Ethyl— amine was distilled into B by cooling B with a dry-ice isopropanol bath. The quantity of ethylamine (EA) used was adjusted so as to form a 0.1 M solution of C. After removal from the vacuum line, the solution of C in ethyl— amine was transferred through the glass frit into compart- ment A. Upon contact with the sodium mirror a dark blue solution formed immediately. The contact with sodium was maintained with shaking for several minutes between 0° and -60°C. The solution was then warmed to 0° to 10°C and allowed to filter through the frit by cooling B. Upon cooling, shiny gold—colored crystals of Na+C ' Na— formed. 56 pcoloxtU .1 QQOO o>_o> E33o0> .Imz . o+mz mo cofluoummmum HON Hommm> .m ousmflm 2.; omeooU BEE 02 e Nessa “*0 _Omm Eanoo> toe cdcgtmcou 57 The excess liquid was again poured into A to insure that all of the sodium which could be dissolved was in solu— tion. Since sodium is insoluble in ethylamine in the absence of C, the solubilization stopped when the reaction 2Na (5) + C + NaC+ + Na7 had used up all of the free complexing agent, C. After the final crystallization process, the super— natant solution was poured into A, the vessel was again connected to a vacuum line and all of the ethylamine was removed by distillation. Purified diethyl ether (purified in a similar way to THF) was distilled into B. The crystals were only very slightly soluble in diethyl ether, which therefore serves as a convenient washing solvent. The very light blue solution formed in diethyl ether was poured into A and then redistilled back into B. This was repeated several times until all traces of white or blue material had been removed from the crystals. The diethyl ether was poured into A and then distilled out of the vessel. Pure crystals of Na+C ° Na- remained in the vessel. The elemental analysis, some properties, and the crystal structure have been reported elsewhere.36.38'52'53 58 4.6 Solution Preparation 4.6.1. Salt Solutions for NMR Exchange Studies Reagent grade NaBr, NaI and Na¢4B were dried and used without further purification. Samples were prepared in a dry box or glove bag under an inert atmosphere of dry nitro- gen or argon. 10 mm O.D. thin-wall high resolution NMR tubes purchased from Wilmad Glass Company were modified by placing a standard ground glass joint at the open end. This provided an air-tight seal. 4.6.2. Preparation of Metal Solutions Figure 10 shows the Pyrex apparatus used in most of the alkali anion NMR studies. Preparation of sodium solu- tions was done by two methods. In the first method, the 2,2,2 cryptand was dropped into the vessel and rested on the glass frit. An excess of sodium metal was sealed into the glass side arm and the entire vessel was evacuated. Sodium was distilled through the side arm and into bulb A to form a shiny mirror. Ethylamine was distilled into B and a method similar to that discussed in 4.5 was used to form a blue solution, which precipitated gold crystals at dry ice temperatures. The crystals could be rinsed off with diethyl ether as discussed previously and then any of the solvents could be distilled into the vessel to form a blue solution which contained Na+C : Na- (%O.l to 0.2 M). The solution was then poured into the bottom of the NMR 59 .22.] . Teflon valve Q I i vacuum C glass 9%3 frit 10 mm NMR tube V B 2,2,2 cryptand .- ‘vA :. sodium metal Figure 10. Vessel for preparation of NMR samples. 60 tube and frozen with liquid nitrogen. The vessel with its frozen solution was pumped to a pressure of «’10-5 mm of Hg and the NMR tube was sealed off at constriction C. In the second method, the solutions were prepared directly by distilling the solvent to be used into the vessel containing sodium and 222 cryptand. The 2,2,2 cryptand was dissolved and a blue solution was formed with the previously discussed method. However, in this case no crystals were isolated, the blue solution was poured into the NMR tube and sealed off under vacuum as before. Cesium and rubidium anion solutions were prepared by the second method. Although no + — + 222 Rb and CsC222 Cs were observed to precipitate in solution, shiny bronze films were observed gold crystals of RbC on the glass walls of rubidium samples and copper colored films were observed on the glass walls of cesium samples. Concentrations of all the solutions were adjusted to approxi— mately 0.1 M in 2,2,2 cryptand. The equilibrium solvent + _ C + excess M 2 (C222M ) + (M ) 222 lies very far to the right for all metals and all solvents used. Sodium solutions were stable at temperatures of +5°C for several days. Rubidium and cesium solutions were stable at dry ice temperatures for a week, but at ice temperatures decomposition occurred within an hour. Thus, the working temperatures of rubidium and cesium solutions were maintained 61 below —30°C where no significant amount of decomposition occurred on the time scale of NMR experiments (3 W 6 hrs.). 4.7 The NMR Spectrometer A pulsed multinuclear NMR spectrometer was used for most of the studies. The design is similar to one reported 127 A block diagram of the spectrometer is by Traficante. shown in Figure 11. The spectrometer consisted of three main parts. The first is a tuned transmitter/receiver sec— tion which operates at 56.44 MHz. The second part consisted of a network of double balanced mixers coupled to a frequency synthesizer. The third part is a wideband transmitter/re- ceiver network coupled to a single coil tunable probe. A magnet from a Varian DA-6O spectrometer was used and main- tained at 14.09 Tesla. With the mixing network, the broad band amplifiers and tunable probe it was possible to observe NMR signals in the range of 5 to 30 MHz. Changing frequen- cies simply required adjustment of the probe, the tunable power amplifier and the frequency synthesizer (which could be done in less than 15 minutes). A key feature of the spectrometer is the external look. In collaboration with David Wright a small single coil probe (lock probe) was constructed and tuned to the proton frequency. The probe was filled with doped water (linewidth approximately 4 Hz) and made to insert into the main NMR probe. The lock probe was connected to the normal 62 a 0 w 0 D 523355.232» $02.52 024.393 02:22 «gaufiimtimzée _ ow23t $5: I? n n 9045: . a _ .23.? _ .52ng fl 1 M £2 on _ _ _ _ . _ _ III _ _ u n «9850 _ _ unfit _ . u u n .525 u . _ out _ _ I _ u u 22 _ _ _ u " 5.022 1. wmdm 6.. u _ £2 an n E m u ._ . $532. _ 134232. _ _ _ . _ 1 _ _ I .(m _ z . 3323 u u :2: t. I _ . _ _ u . _ new” I“ 3 n _I in 4. . a . 7.ka T . _ _ _ _+ 63 proton lock circuitry of an unused Varian DA-60 spectrometer. In this manner it was possible to lock the magnetic field on a proton signal and perform magnetic resonance experiments in the frequency range of 5 to 30 MHz. The maximum drift in an eight hour period was less than 6 Hz. The details of the external lock probe construction are given in David Wright's Ph.D. thesis.128 The spectrometer itself is controlled by a Nicolet 1080 computer system equipped with a Diablo magnetic disk. The hard-wired interface and software are described elsewhere.128 4.8 Temperature Control and Calibration Temperature regulation was accomplished by using a standard Varian heater-sensor and V-4343 variable tempera- ture controller. Temperatures were measured with a cali— brated Doric digital thermocouple. A thermocouple was placed in a fixed position inside the probe, at a height which is just below the bottom of a sample tube. To make sure that the temperature at the thermocouple and the tem— perature at the sample tube were the same, a calibration procedure was used. A mercury thermometer and a thermo— couple were placed inside the bottom of an NMR tube with 1 ml of solvent. The thermometer bulb and thermocouple were positioned at the center of the receiver coil of the NMR probe. The temperature was varied from 0° to 100°C and the thermometer reading along with the readings from 64 the thermocouples inside and outside the sample tube were recorded. The results are given in Table IV. From these results, it was not felt necessary to perform any tempera- ture correction and all thermocouple readings were used as read. 4.9 Data Reduction Free induction decays were stored on magnetic disk. The data were Fourier-transformed without applying smooth— ing or exponential weighting. The transformed data were zero-order phase corrected and punched onto paper tape with a modified version of RELAX 2 (a program writtEn by David Wright).128 The modification to RELAX 2 is given in Appendix A. The data from paper tape were punched on keypunch cards in octal format. The octal formatted data were converted to decimal format by program CONVERT (Appendix B). The decimal formatted data were analyzed by a generalized I“ weighted non-linear least-squares program (KINFIT).129 65 Table IV. Temperature calibration for NMR. Thermometer Inside Thermocouple Outside Thermocouple (°C) (°C) (°C) +4.1 +2.9 +2.9 8.5 7.6 7.9 13.2 12.3 12.3 18.8 18.2 18.3 26.0 26.0 25.9 24.1 23.7 23.7 31.9 32.0 32.1 37.9 37.7 37.7 47.2 47.0 46.9 51.9 51.7 51.5 57.0 56.9 56.6 61.7 61.6 61.4 66.4 66.3 66.2 73.4 73.4 73.2 78.4 78.6 78.3 86.5 86.7 86.3 93.9 94.3 93.9 98.9 99.4 99.1 V. SODIUM-23 NMR STUDY OF SODIUM ION - SODIUM CRYPTATE EXCHANGE RATES IN VARIOUS SOLVENTS 5.1 Introduction Macrocyclic polyethers (crown ethers), first synthesized by Pederson,32 and macroheterobicyclic diamines (cryptands) 34 firSt synthesized by Lehn 2; al. form very stable complexes with alkali metal and alkaline earth metal cations. X-ray structure determination of cryptate complexes formed from 272:2 cryptand (C see Figure l) and alkali metal salts 222' show that the alkali cation is contained within the central cavity of the ligand.l30’131 Association constants for complex formation between many of the synthetic complexing agents of both the crown and cryptand classes and a variety of metal 132-134 salts are known for aqueous solutions. Except for association constants in methanol, very few others have been determined in nonaqueous solvents.l32-134 Even fewer rate data exist for the exchange of metal ions between solvated and bound sites. Exchange rates at the coales- cence temperature have been reported from PMR studies on D20 solutions which contain C222 and half the stoichiometric 39 amount of alkali metal salt. From these results it was clear that sodium cation exchange might be slow enough to 23 be observed directly by using Na NMR techniques. Ceraso 40 investigated the kinetics of complexation reactions of Na+C222 cryptates in EDA solutions by using 23Na NMR and Dye 66 67 techniques. Their work is the first example of sodium cation exchange which is slow enough to exhibit clearly defined separate signals for two environments. Cahen 33 al.135 studied the Li+—cryptate exchange kinetics in water and in several nonaqueous solvents by using 7Li NMR. Cryptands smaller than C222 were used. No other quantitative informa— tion exists for alkali cation exchange rates in the presence 136-138 of C Schori e: 21. studied the kinetics of com— 222' plexation of crown ethers with sodium and potassium cations 23 39K NMR lineshape anal- in several solvents by using Na and ysis. Since for the crown case the chemical shift between the free and bound site is negligible compared to the 23Na and 39K linewidth of the bound site, their experiments only exhibited one apparent resonance. Under their conditions, it was possible to reduce the complete lineshape expression for two site chemical exchange to a single Lorentzian line shape function. This is a special case of the treatment used here. In this chapter an NMR lineshape analysis of the tempera- ture dependence of sodium ion exchange in the presence of C222 in H20, EDA, THF and PYR solutions is given. Lineshapes are calculated from the exact expression for a general two site exchange of uncoupled spins and are fitted to the ex- perimentally observed lineshape with the aid of the generalized weighted non—linear least—squares program KINFIT.129 23 We have chosen to use the Na pulsed Fourier transform NMR technique because the advantages over continuous wave 68 NMR techniques are threefold: 1. No modulation distortion from sidebands occurs (particularly a problem with sodium, due to its range of relaxation times); 2. No saturation broadening occurs; and 3. Signal averaging is accomplished in a much shorter time period. Despite the low sensitivity of 23Na NMR compared to that of PMR, its use has two advantages over PMR: l. The chemical shift range of sodium nuclei is much greater than that of protons; and 2. Deuterated solvents are not required. 5.2 Determination and Interpretation of the Line Shapes The bloch equations which describe the motion of the X and Y components of magnetization in the rotating frame, when modified to include chemical exchange, are given byl39’140 E— + GAGA = ‘lYHlMOA + TB GB — TA GA 5. 1 —Ci-‘E— + (XBGB : —1YH1MOB + TA GA "' TB GB , 5 . 2 where G = u + iv G = u + iv a - T_l~i(w m) A A A’ B B B' A 2A A ’ 69 0‘B 28 B frequency field, MOA B of magnetization for sites A and B, and TA and TB are the = T -i(w -w), w is the variable frequency, H1 is the radio and MO are the equilibrium 2 components mean lifetimes in sites A and B respectively. Other symbols have their usual meaning. The solution to these equations dGA dGB for slow passage conditions (?fi7 = TEE = 0) and for transient conditions (Hl = 0) have been shown to form a Fourier trans- 141,142 form pair. The solution for slow passage conditions is given in Appendix C. The shape function is given by G(w) = YHlMoiIcose - Rsin0} 5.3 I = 82+T¥; R = Hg:§%- 5.4 S +T S +T P P A B T S = ——— + ——— +-—————— - T(w -m)(w -m) 5-5 T2A T2B TZATZB A B P P U = 1 + T(TB + T—é- 5.6 2A 2B w -m w -w A B T = P w + P w - w + T[ + ] 5.7 A A B B TZB TZA V = T[PBmA + PAwB-w] 5.8 T T T = A+ B 5.9 T T A B p , p = 5.10 A TA+TB B TA+TB 70 where I represents the absorption mode lineshape and R represents the dispersion mode lineshape. wA,mB and T2A’ T2B are the Larmor frequencies and transverse relaxation times of sites A and B in the absence of exchange. PA and PB are the relative pOpulations in sites A and B. Equation 5.3 predicts the lineshape throughout the entire range of ex- change from the extreme slow limit to the extreme fast limit (slow exchange occurs for T >> ‘ 1' ), fast exchange occurs w -w A B for T << ——¥L——-). The exchange time I together with the (MA-MB) relative pOpulations PA and PB contain all of the kinetics information. 5.2.1. Measurements in the Absence of Exchange In order to evaluate exchange times through application of Equation 5.3, we must obtain information about the Larmor frequencies and transverse relaxation times for sites A and B in the absence of exchange (A denotes a solvated sodium cation and B denotes a cryptated sodium cation). In addi— tion, the populations P and P A for exchanging systems must BI also be known. In many studies of systems which undergo two 1 d 13 site chemical exchange, H an C NMR lineshape analysis is used to obtain T values. In some of these studies the assump- tion that l/T2A = l/T2B = 0 is made. This assumption is not justified for sodium since its relaxation rates are usually much larger (broader lines) than those of 1H or 13C and these relaxations contribute significantly to the observed NMR lineshapes. 71 Since the equilibrium constant (Ka) for complex forma- tion between Na+ and C222 is large (Ka > 103) in all sol- vents and at all temperatures, PA and P are directly deter- B mined from the mole ratios of C222 and sodium salt added to solution. To obtain MA,wB, T2A and TZB’ two solutions in each solvent were prepared. The first contained a sodium salt with no C222 and the second contained a sodium salt and equimolar C Because Ka is large, the second solution 222' contains only bound sodium cations and these ions cannot undergo chemical exchange since the concentration of unbound 23Na NMR of each of these solutions sites is too low. The was studied as a function of temperature. All experimental lineshapes were found to be Lorentzian. The values of w and T2 were determined by fitting a Lorentzian line to the observed signal. Complete character- ization of the intensity of a Lorentzian signal by the Fourier transform NMR technique requires determination of six param— eters. These are the amplitude, K, the Larmor frequency, w, the linewidth parameter, T2, the height of the baseline, C, the zero—order phase correction, 00, and the first-order phase correction, 0'. The zero—order phase correction is constant and independent of frequency. It determines the contribution of dispersion and absorption mode signals to the observed lineshape. Electronic filtering and the finite length of a 90° pulse that initiates a free induction decay, as well as delays in the start of data acquisition can all 72 lead to a first-order phase correction. The first-order phase correction varies linearly with frequency (and hence channel number, j) according to 0' = 01 ° % in which N is the total number of channels in the spectrum and 01 is the total change in phase over the entire spectrum. In terms of these parameters, the intensity is given as a function of frequency by K’T GIN) = 2 2 2 {cos(00+0') — T2(IIIO-III)Sin(90+9') + C 1+T2(wO—m) 5.11 Commonly the phase parameters and the baseline height are instrumentally adjusted by visual inspection of the signal to obtain a symmetric absorption—mode signal with a zero baseline. The value of 01 is constant for given instrument settings and may be accurately determined by measuring the signal from a reference sample at several magnetic field settings which span the entire frequency range to be used. Once 01- is evaluated in this way it need not be changed from sample to sample and is not an adjust- able parameter. By contrast 00 is influenced, not only by instrument settings, but also by sample position, tempera— ture, etc., and therefore it must be adjusted for each sample. Since in any event, the parameters must be adjusted (either numerically or instrumentally), we have chosen to evaluate all parameters except 01 by fitting the Lorentzian Equation (5.11) to the observed spectrum by a non—linear 73 least—squares procedure. A typical example of a fit for 0.6 M NaBr in EDA solution at 26.3°C is shown in Figure 12. The variation in linewidths and chemical shifts with tem— perature for free and unbound sodium in each of the four solvents is listed in Tables V-XII. The measurements of T2 and w for free Na+ in H20 and PYR were made visually. The data for each solvent are plotted in Figures 13-16, (semi- 1og plot of l/T2 ys 103/T). Simple exponential behavior of the variation of the relaxation rate with temperature is observed for THF, EDA and H20 solutions but not for PYR solu— tions. As shown in Figure 16, the relaxation rate levels off between 90 and 140°C. This deviation from exponential behavior might be caused by changes in ion-pairing with temperature. In all cases the activation energy (see Chap— ter III) is smaller for free sodium cations than for bound sodium cations in the same solvent. All chemical shifts are referenced to the infinitely dilute aqueous sodium cation (a positive shift corresponds to a high field shift). The Larmor frequencies show only a slight dependence upon temperature. The value of m at any temperature between those examined was determined by graphical interpolation. A sample of saturated aqueous sodium chloride at 25°C was used as an external reference for both chemical shift and linewidth calibrations. (The chemical shift of saturated aqueous sodium chloride relative to infinitely dilute aqueous sodium ion is —0.7 ppm). Linewidth contributions from magnetic field inhomogeneties ranged from 2 to 6 Hz and 74 .uoad ecu mo coehsaomou ecu segues meow on» one guess ucHOQ cmumasoamo cam kucwefiummxm no .N .Dcflom omudeoamo m .o .ucflom Hmucmseummxm cm mucmmmummu x .«om CH Hmmz z om.o maecemucoo coHDDHOm o How mmmcmmcfla MZMN ecu mo memwamcm BHEZHM Hecammu m .NH ouomflm roLx \Cu..( C otooooooooooooooooo lmIIIIuIIIIUIIIImIIIIEIIIIUIIIIMIIIIwIIIIuIIIIvIIII.IIIImIIIIEIIIIx IIIIIIIIIIIIIII u IIIII .IIIIr IIIIIIIII _ UH nun nuuu nun" ”Guy. run IInu uqu unun unuuv H x C nun" nnx nun n... . — CM flu V" I H w H” H: . H "H u. a U H «N u H n U. M p H a l n n e x n n U C u » H H u n w H a _ n U H C u p x h p . P . k , W H H U H H w . l L U C U a w M w u e L M . . J J A .. .. u M i l U n u a F H w » w p i .u u a u . F "4 _ a . iJIIIIJIIIIUIIIIrIIIIJIIIIJIIIIrIIIIJIIIIJ IIIII . IIIII . IIIII IIIIrIIII III- IIIIuIIIIJIIoIt IIIIIIIII rIIII. 75 Table V. Variation of the relaxation rate and chemical shift of free sodium cation with temperature in ethylenediamine.a T(°C) 103/T(°K) 1/T2B(sec)b 6(pmeC 20.6 3.404 255.3 —13.86 26.3 3.340 220.3 —l3.75 32.1 3.276 192.8 -l3.67 38.0 3.214 173.3 -l3.65 43.5 3.158 154.7 -13.61 49.1 3.103 139.7 —l3.59 54.9 3.048 124.0 -l3.56 60.0 3.002 113.8 -l3.53 64.5 2.962 104.8 -13.47 70.1 2.913 94.79 -l3.42 a) 0.6 M NaBr in EDA solution. b) Natural relaxation rate. c) Referenced to infinite dilute aqueous Na+. 76 Table VI. Variation of the relaxation rate and chemical shift of bound sodium cation with temperature in ethylenediamine.a T(°C) 103/T(°K) 1/TZB(sec)b 6(PPm)C 20.2 3.409 273.6 10.41 26.6 3.336 231.5 10.65 32.3 3.274 202.8 10.90 35.1 3.244 182.6 10.83 41.4 3.179 158.0 10.92 48.3 3.111 138.3 11.21 52.4 3.072 127.0 11.10 59.0 3.011 110.4 11.17 67.7 2.934 95.79 11.24 a) b) C) 0.3 M Na+C - Br— Natural relaxation rate. Referenced to infinite dilute aqueous Na+. in EDA solution. 77 Table VII. Variation of relaxation rate and chemical shift of free sodium cation with temperature in water.a T(°C) 103/T(°K) l/T2B(sec)b 6(ppm)C 7.8 3.556 37.5 0.3 13.9 3.484 27.3 0.3 21.2 3.397 24.9 0.3 33.3 3.263 18.6 —-— 36.2 3.233 17.1 0.6 a) 0.2 M NaI in H20 solution. b) Natural relaxation rate. c) Referenced to infinite dilute aqueous Na+ 78 Variation of relaxation rate and chemical shift of bound sodium cation with temperature in water.a Table VI I I . T(°C) 103/T(°K) l/T2B(sec)b 5(ppm)c 2.8 3.624 520.7 7.45 8.3 3.553 438.1 7.68 13.4 3.490 355.3 7.84 19.2 3.421 311.3 8.12 21.0 3.400 282.3 8.45 25.0 3.354 245.5 8.54 29.7 3.302 233.1 8.53 35.3 3.242 184.3 8.83 40.6 3.187 164.5 9.07 45.9 3.134 149.3 9.28 50.7 3.088 134.3 9.46 54.6 3.051 122.4 9.54 60.8 2.995 110.5 9.72 66.1 2.948 100.5 9.86 a) b) C) 0.2 M Na+C - 1' Natural relaxation rate. in H20 Solution. Referenced to infinite dilute aqueous Na+ 79 Table IX. Variation of the relaxation rate (1/T2) and chemical shift of free sodium cation with temperature in tetrahydrofuran.a o 3 o b C T( C) 10 /T( K) l/T2B(sec) 5(ppm) 25.7 3.346 96.92 +7.43 30.7 3.291 94.29 +7.53 35.4 3.241 91.97 +7.58 42.8 3.165 88.60 +7.72 47.9 3.114 87.43 +7.82 53.1 3.065 85.79 +7.93 a) 0.4 M Na¢4B in THF solution. b) Natural relaxation rate. c) Referenced to infinite dilute aqueous Na+. 80 Table x, Variation of relaxation rate and chemical shift of bound sodium cation with temperature in tetrahydrofuran.a T(°C)I 103/T(°K) l/T2B(sec)b 6(ppm)C 19.7 3.415 169.5 11.89 25.1 3.353 154.3 12.03 30.3 3.295 139.0 12.17 35.4 3.241 129.5 12.26 42.1 3.172 119.7 12.52 46.6 3.127 111.1 12.48 53.2 3.064 101.4 12.62 59.5 3.006 94.72 12.57 64.2 2.964 89.25 12.67 68.4 2.928 86.46 12.74 70.9 2.907 84.57 12.81 a) 0.2 M Na+C ° ¢ B— in THF solution. 4 b) Natural relaxation rate. c) Referenced to infinite dilute aqueous Na+. 81 Table XI. Variation of relaxation rate and chemical shift of free sodium cation with temperature in pyridine.a T(°C) 103/T(°K) l/T2B(sec)b 6(ppm)C 22.6 3.383 86.8 —1.2 45.0 3.145 70.4 -0.6 57.6 3.025 62.1 -0.3 69.0 2.924 57.2 —o.2 80.3 2.830 53.9 0.0 88.3 2.768 55.5 0.1 100.8 2.675 53.9 0.3 112.4 2.595 53.6 0.4 125.4 2.510 53.9 1.0 133.8 2.458 58.8 1.1 140.3 2.420 57.2 1.3 144.8 2.393 58.8 1.3 a) 0.2 M Na¢4B in PYR solution. b) Natural relaxation rate. c) Referenced to infinite dilute aqueous Na+. 82 Table XII. Variation of relaxation rate and chemical shift of bound sodium cation with temperature in pyridine.a T(°C) 103/T(°K) l/TZB(sec)b 6(ppm)° 25.5 3.348 150.9 12.38 32.9 3.267 134.8 12.54 38.6 3.208 124.9 12.56 43.7 3.156 116.5 12.65 55.1 3.047 97.43 12.76 61.4 2.989 90.14 12.82 68.8 2.924 83.87 12.93 74.9 2.873 78.12 12.98 82.6 2.811 72.75 13.05 88.6 2.764 67.45 13.10 102.2 2.664 63.73 13.22 109.2 2.615 59.06 12.79 114.0 2.583 61.01 13.28 121.2 2.537 60.24 13.18 128.6 2.489 60.50 13.19 133.8 2.457 57.71 13.18 140.2 2.419 59.33 13.07 + - . . a) 0.2 M Na C ° ¢4B in PYR solution. b) Natural relaxation rate. c) Referenced to infinite dilute aqueous Na+ 83 .8 . musummmmsmp .aom as loo NNNO oz 2 m.o cam ADV ummm z v.0 mchHMDGOD mcoHDSHOm “Om .me enemas macaombm Hmooudwomu m> 62mm new NE\H mo muoam modesmm r.x...§~o F to a.» as a s _ _ _ ow \O 01 00—. 03% .\.\ .4 .h an \OMN Ioou m \0 (u \m\ \\ Icon 84 500- Figure 14. 3.2 3.4 3.6 103/T(°K") Semilog plots of l/T2 for 23Na ys reciprocal absolute temperature for solutions containing 0.2 M NaCl (0) and 0.2 M Na+c222 . Cl“ (0) in H20. 85 Imee . NN~o+mz z ~.o 6:6 10V ousumuomfimu DDSHOmnm Hmooumfloou m> 62 .53. ca AOV meemz z e.o mcHCHmucoo mcoHDDHOm mom Hoe ~e\e mo muoea modesmm .me magmas mm A Rosina. a.” _..n QN _ e _ _ _ Ayn Illfivlllllgvlllg'.ll;'.||l||\\kU\\AVAU\\\I1 III.) 0 o \o 100.. \ o\o .3 o\\ )I no\\\\ “a o\ o \o {w ICON 86 .msm an loo Imve . mmmo+mz 2 m.o cam ADV meemm z m.o ocecamucoo mcowusaom Mow ousuouomfiou musHomnm Hmooumflomu m> 62mm How NB\H mo muoam moHHEmm .oa mucosa 1.61.5.2 «.6 ad TN _ l l _ l s cc 0 O O O o oo o . .,.... O 0 \Sn. 0 109 w 0 7w 0 o o OON 87 were, for all cases except aqueous solutions, a small frac- tion of the measured width. Inhomogeneous line broadening varied from one set of experiments to the next. This con- tribution to the linewidth was determined for each set of experiments by using an aqueous reference sample with a known relaxation time (saturated aqueous NaCl; natural line— width is 8.0 Hz at 25°C), according to (l/Tz) (l/Tz) (l/Tz) obs. nat. inhomo. This value was then added to the previously determined value Of (1/T2)nat. for each of the sites so that the values of l/T2A and l/TZB used in the following line shape equations include both the natural contribution to the linewidth and the inhomogeneous contribution. This method assumes that the shape function of the field inhomogeneity also has a Lorentz form, which is often not true. However, for all cases but free sodium cation in H O, the inhomogeneous con- 2 tributions to the observed linewidth are small and are ap- proximated well by a Lorentz shape function (see Figure 12). Even the narrow lines observed for free Na+ in H20 were Loren- tzian. Values for (1/T2) at any temperature were evaluated by fitting the following exponential equation to the observed data (corrected for inhomogeneity). eEr/R (l/T-1/298.15) (l/T nat. ) = (l/T ) 2A Or B nat' 2298.15 88 This was done for all solutions except free sodium in H20 and free and bound sodium in PYR. In these cases graphical extrapolation and interpolation was used. The values for Er are listed in Table XIII. 5.2.2. Evaluation of Exchange Times To evaluate the exchange time T, Equations 5.3, 5.5, 5.7 and 5.8 were modified as follows G(w) = K{Icos(eo+e') - Rsin(00+e')} + C 5.3' P P A B T S = ——— + ——— +-—————— - T(w +A—w)(w +A—w) 5.5' T2A T2B TZATZB A B wA+A-w wB+A-w T = P w + P w + A - m + T[ A A B B T2B T2A V = [PBwA + PAwB + A - w]. 5.8' Application of Equation 5.3' requires evaluation of six parameters. These are amplitude, K, baseline height, C, zero-order phase correction, 00, first-order phase correc— tion, 0', frequency shift, A, and the exchange time, T. Again the first-order phase correction was not adjusted num— erically, but was measured visually for a given set of in- strumental settings as described previously. Thus, in most cases, we chose to adjust five parameters to obtain the best fit of Equation 5.3' to the observed data. The frequency shift parameter was introduced because of experimental Table XIII 89 Activation energy, Er in various solvents. , for solvent reorganization Concentration Er Ion (M) Solvent (Kcal) Na+ 0.6M NaBr EDA 3.6 Na+C 0 3M N +C 'B I EDA 4 4 Na+ 0.4M Na¢4B THF 0.86 + + — Na C222 0.2M Na C222 94B THF 2.7 + + — Na 0.2M Na C222 I H20 4.8 90 difficulties with referencing absolute frequencies from one set of data to the next. The relative chemical shifts at the two sites were not adjusted. The errors in precise referencing of one set of data to another were caused by the method of external referencing. If an internal reference in each sample and an internal lock had been employed, there would be no need for the correction term A. The frequency shift parameter allows the absolute frequency in Equation 5.3' to shift, without changing the shape of the function. In most cases A was less than 0.5 ppm. For each spectrum analyzed, 90 to 99 data points were found to be more than sufficient to determine T. The program used (KINFIT) gave complete statistical information about the fit to the data, including standard deviation estimates for each of the parameters and the multiple correlation co— efficient, which gives a measure of the coupling of each parameter to all of the others. Coupling between T and the other four parameters was lowest at intermediate rates of exchange (I e l/(w This is expected, since I has its greatest influence upon the lineshape in the intermediate region of exchange. 5.2.2.1. Exchange Times in EDA Solutions — Three solu- tions (EDA l, 2, and 3) each containing 0.6 M NaBr but with variable amounts of C222 (0.15, 0.30 and 0.45 M respectively) were examined. Selected spectra for EDA l, 2 and 3 are shown in Figures 17-19. Spectra from solutions EDA 1 and EDA 3 91 Figure 17. Spectra at various temperatures for a solution of 0.15 M C222 and 0.6 M NaBr in EDA (1 ppm = 15.87 Hz). 92 1 76 . 40 I I g .I ‘f\ n r 'I I I \ 7 :\ ,I I \. I l I /’ \ ‘ / '/ I f' | )\ l/ I \ ) I ,I I {I' \ / I’ I, I1 I“ f f I 1,, 1‘ I‘ ll / \fi I f. / 'I I I i f v/ _ I a k I L O /« I w I w. 5 0 . 2 P J, 1 t \ I / I II I A , / J I I I l I / I I I. I I f I. III 1 I III. ' / I I \ I If '/ I \ I‘ I / I / ' I ‘ ’ I II I. I t . J (I k \ / . I I I I \. I I I l \ I ) I I I ‘. I 4 2 . 4 ° ‘ I \\ I I I I“ L \x" 'I II, I . \‘I \\I . /// II \ / I )\ / f‘j / \I \\ r l/ ‘1 I ’ J I T f f I I ' ’ /' \ I 21.30 Figure 18, Spectra at various temperatures for a solution of 0.30 M C2 2 and 0.6 M NaBr in EDA (1 ppm = 15.87 Hz). 10 ppm H—~—+> 70.2O 58.8O Figure 19. Spectra at various temperatures for a solution of 0.45 M C2 and 0.6 M NaBr in EDA (1 ppm = 15.§% Hz). 94 were analyzed by using Equation 5.3'. The spectra from EDA 2 were also analyzed by using Equation 5.3', but the value of 61 was determined by computer adjustment of a modified form of Equation 5.3'. The A term was dropped and ”A' wB and 91 were allowed to float (along with K, C, T, 60) to obtain the best fit of Equation 5.3' to the experimental data. The three lowest temperature spectra were fitted and the average value of 61 was computed. This value (61 = 264°) was then used to analyze all the spectra from solution EDA 2 by using the unmodified form of Equation 5.3'. Some typical fits for these data are shown in Figures 20-25, Variations in T values with temperature are listed in Tables XIV-XVI. 5.2.2.2. Exchange Times in THF, H20 and PYR Solutions - A single solution of each solvent was examined. Equation 5.3' was used and no 91 corrections were made (61 = O). For the H20 solution, the A parameter was not used. Some of the fits for H20 solutions are shown in Figures 26 and 27. Noticeable intensity differences between calculated and observed values at low temperatures in the vicinity of the broad peak (bound cation) occurs with the experimental intensity lower than the calculated intensity. This may be partly caused by the spectrometer delay time. After the pulse, a delay time of 200 to 300 microseconds occurs before collection of the free induction decay. The area under a peak in the frequency domain is proportial to the intensity of the free induction decay at time zero. If a 95 I J'I—o—n—n flan—o.— n-ul—‘fin-a...‘ ‘ o 4 I l O l l g l I . I a s a I 'I u ( - I , ' U I u I 1 - '. I: n I: ‘ :: xl-I: 0 '35’0- _ l I: J I ’ . ’ ‘ s: I. :0 ":3 1:21: I Ir:===:v =:::.—nrm ‘::::-"J?::;::nn [(x:::::r:::r:::-:: :c --------S---—§---—§----Sl I l I I 00 I I ‘ ' . o I I z I I I 9 X ‘I I 0 I I I a I . . 0 I 2 L I I a I = c. 'y I I I ' I I I i . 1 I i I I I I I I ’- I I s s I t I I 0 I l = I l I 1 = 0 ‘5 I = I I 0 cl l I A n: l I 1 3:1 I I - 0.! Q I = 0:1 I I : r)::: I I 1 ::::;n1 I 1 . x-::- -r~:on=n 1 _ a-,_:;: 1 ll l:::'.-:-_-.::: 2‘, I- - .'--- .-‘-—-. ‘- — v--"_,---—‘.-'--'.-~--"‘ -".----‘,—---‘.----‘)—---§----ry----‘n----';----S--‘-S-"-'II O'Cufl'IOI‘JODVOIO‘ 9 ( .l”[l).\ o .ll‘u (pyJI (.A.Q :-.-—§----‘.----‘~---".----<-~—-§----‘~----§----1---.-H-~--§----‘.~-o-§----g----S“--‘v---’S‘---‘~----S°'--‘II (I I ” I I Q I. I l I O I I I < q. I I I I < I, I . < I 'I I I I l I b S c. I n I n I I I r. J, I . . I I I ‘o I I, I ' I I I I u I 0 I‘ u; g on 'I I n I I A l 1:! I L 0: I :00 t ' 11:3 ‘5 I n“ 3: [ :{ ::;=l I :1-: 3'::O I l I "z I l.’:;:::::n; k ---- 'H _‘.. I :::::; :‘y I - a? -'.-e - r,,‘.,___._-_-_r'---_-,--_-<.--__g--__q-,_,g_---c.-_-_q-__-¢.- --“----§----C~----'§I .0000. an o-1oou I r ,LlfOfl o .IH‘I CHY’I (31.9 Figure 21. Computer fit of spectra obtained with 0.15 M C222 and 0.6 M NaBr in EDA. (a) 64.9OC; (b) 53.9Oc. 97 ----s----s----s----s----s----S----s--—-s--‘-s‘---s-—-.s----<----s----s----s----s----' .--—-<----s-~--n| I - I 1 l l I '00 l .00. S I 00 I I I O l o I ‘ O I n ' 4 0 ' n l ' 0011 o I ' ' ll In I ' 0 u I I I ‘ no I 0“ l I OX 0 g I 00: I no ‘ 0 l 000W)= 0 l x . - I I o Q . ‘ I 0 l a 'I o ‘ I u z! 9 ° I I o u o 1‘ x ‘0 I ll ‘ ') :0 0 I x I s , 1 I ' =2 0 I = s a ” . ' w °nn I n . I I, O 0 I I n 0 I ‘. lemma ’1 ‘ mm 1 1 ¥ On 1 l ~(m x 1 III“ 0| I X l ‘1. __--5---_g____r'._-_r_-_-_¢’ ____________ I, __________ n: r """"""""'°° C run g- a: I S """'g"“‘ “““ ‘ ---- ~----‘ ----- S--—-s---—§_---HI ----5----s----s-~--s----s----s----a----S----s----s-—--n-—--a——--n_---n----«----nn--»n--—-n—-—-s—---n: I I I no: I I ‘ o - on a U) I x 01 l l O = X I 0 I 1. O I t‘ x x I 0 O I l I 0 - I I| l ‘ S l n I l 0 x l 1 I O n I x l I S I ‘i I 0 = 0 I I 0 I I x x x l x l b 0 0 0 X ‘x I 1 l O I l 0 ~ 0 I x x I x l o o = I x l I n I! O O l x I t O t ‘ 1. x n l I " OX 01 | 0: Ir: 0 I z m x 0 ‘ I ‘> ' n-U l n: R o orrn I n1 : I Cry 0 I O l' I ‘, ::K 0' " om um I 7’11: In” I I ~ I' x n 0 II I; ’) I III S! l . l r ‘I - l‘- -‘1 1 r ___1 I‘_>_ 1_____l____’____r'-___(l___>r’ _ T|7_A_' _____ ,_-__l‘_‘__ ‘-__7( ..... "__ II ”unanauuxrg Figure 22. Computer fit of spectra obtained with 0.30 M c222 and 0.6 M NaBr in EDA. (a) 45.00c; (b) 30.40C. 98 S—---s——--s ‘. s s s = ‘ r. 5 St I I o: I 0 ' I X l n . i, n - I ‘ I l o ' I I I n o ‘ l I 6 I - I 0 I . l i a . - g I n I - I x O o ? x x I . l l . o l I x l s I x 9 - l l . =1 7 n 0' =; I :1 ' I u n: I or I .7 s l :01 I “an xo::= I I , r,‘ ‘ zgno l l l I I, , , le aura-.4 x I ‘1 071:» ,-_ :g “sq . , ”is. -—s-—-—s—-——r 7 I-»——sA-—-s—~-.n~.~r“us-”J...” ”us. [LA It. I~—--$~-r‘;-—--‘- , s“ <5", 5~7-«mash—AS»e—«s—n-n----‘.-~-—‘.----‘.»—-—‘.—---‘.~«~s----<.»—~-s~-—‘.I x i l l mu) 1’ l l x 0')qu I l l 'n[)'-'_)fl,n x l n I ’v .0 no ‘ I 10") l [O I I 0 I x x l l 10 x I S In 1 ‘~ I ‘ ° I 0 l I n -x I 1 l O! r_ I 0 I I 0 . z s . <. l n I I o x! l b I 0 I 01 l I s o N I m I l I I l I I J . 0! | I Ox S I " I O [)1 | xx 0 l 0 *X I o I I n I I I m 00 I I ex 1 I l X 0 I 8 In In a l l ,1 ::| | I) o I l n I. n I J In! 71‘ I , 1m 0 v \ I I l m A I. m. . I I‘ l I], i Ix” , I, _ . V__L _________________ .'___Vr,____ ,. ,.,_ __r __‘<‘ __.r,_ _,I,.._.t. . ‘,_._.r ._.I,_ _. l -..............-...... (“A _ 0 Figure 23. Computer fit of spectra obtained with 0.30 M 3%? 2 and 0.6 M NaBr in EDA. 50.2Oc. (a) 76.4OC; 99 I-—--§-—-c-----'.----§~---".----H----<----§-c--§----‘.---—§---—'\----‘w----‘x-—--‘.----'.----S----§----')--—-"I 'I I I) . I 0 I . I O . I I l I I a ' I I ‘ o 9 I I = I I ‘y a ‘1 l I I I ‘ I a 'I n '5 I I I x I I n X I 0 I C 'N 0 I X I I = I I <, I : ‘. O I 0 I I :u o I 1-: I l ‘ luxc~7 0; I. x-nmm 0 I 11:0 2 I I'm . = I I") : I I 11V) :: (n 1") 27 I HIV-n r“. ‘ 'm‘nt 3‘ n‘(\ I 1 fI-R'V')» It I )7 I‘I ‘. 'I 2: I! v \'. .____ I,---'-_r_,---_r"-____’<,,-(___-q_--_L.,-_,23__._II-_,_r'____(_‘_,,_r’__,_v_-_-_r"--,1'_,_.,I.---1I-_-_f._-. 1,] 0°000¢voa~owuo¢oc o f ,I.‘,u ('I'V'JI I4 [HA I0.*. 1_-_-' _-_,---_ _-i-_ ._ __—..,_-')____(‘_-_‘S_‘__j___-()____(‘___-1‘_>-(I____(‘___-(.-___H__-_l‘____f_-__C_____ ‘I I I I 0 I I 1 l I z n I I ‘ I I I I I I 2 v Q : H I I I I < r, l I 0 ’ I I I t r, I I : : I b I c, g I I l‘ I I ‘ ” I 0 r, r I I I t I I = I I I C r ' I = l I I I I n 0 I 1 n: I l I '3 l')-II‘)L: = n " I "I : = ’ I 1') =2 t 3 I I I) :_-0( :r: : I I At): a: I c, I. r ‘. I , -:t I I I 0' 0 I I I I (”In H I I ,3): [Iv ‘-- -"'t\I'|"k1I l ‘1 III'I I' I-.—.,-——-‘,--- .--i_._,m A--..-I"- -.--—w.-- _s—- «3. .. r --- - -'.---— ----v.--- - - "--— ' - . .- -’.| .goto‘ococoouoaoono o I .‘.‘,‘v («HUI I'd IIIA I‘!.I Figure 24. Computer fit of spectra obtained with 0.45 M c2 2 and 0.6 M NaBr in EDA. (a) 30.6Oc; (b? 19.30c. 100 ._--$—-—-fi-~--R—-—-§--—-R_---R----R~—~-§~--—§---nR----fl----§---~Sn---§---~R—-~~§----§--‘-9-~~-§~---5I no I I lo I l l ‘5 0 I l ‘ I O l I I 3 S . I «I a I!» l X (I 1 n It I I I x l I n O a ‘v a I ‘ I 0 0 I s I I» ‘ I 0 l I ‘ 0 I n I X s t 0 I I l I l) O ' K I r) (I : ‘ ‘\ I x: m I I :0 7 I I - :x l l (H I n, n ( 0:! \ I -- n: I I I ”I I 1. I; X: -;n I I ‘7: 2‘ ‘ I 1 - .m» mmnv w! I S- :t ll ll(l)lII* 2t, I‘-"k’ ........ '- ._', -_,___,L,_,__'_----,,-,,'.--_-5----t,----l,--__€,-_--l,_--_‘..-__I_- -‘._._-t‘,_., _ g: 9.0.0coanronvcwg.sno r .45,” .,1'.‘I IN [\‘IA «(,4 I~-—-‘D--—-r.—‘-~( .‘_'__\__-_(‘__<_C',__-I.-~--I'____(‘____\'__--r‘__‘-(|--'_IT'___,(._-__"___,l\-___Q‘ I I 0 I I (U!) I I x x I ') K I 0 I I x D I I l I I I s a I r I I = ‘ I I I r. : R I I l I I : I I x I 3 n '3 I r I I I 0 I I b l I I r_ K I = I I (x) l ‘I x j. I 0 l I I 0 I I = ' I ‘ I I 0 O R I z I I l l l I (J O l I I . x I ‘» 1m 0 .' I 1m: - I l x.” l - I _ 1:0 0‘ . I :0 O I (I l I I ‘I II : I I II I I n r (I I I I .'.’I'p'l‘) -- I -,| I \ \ II‘ ' 'Ifilqu l'lI'III I I- ,---- ..-.~.---_-,- ._'- .r.. -I. -.-I.--”k.-.--r.----l\_---'._-_-\,---R-_--‘.-_-_'-_ J. -- ..-'v coocaflanut ::::: orOI‘ ( .I.-,-vvi.v[ I‘I'A I'«.I Figure 25. Computer fit of spectra obtained with 0.45 M and 0.6 M NaBr in EDA. (a) 41.9OC; C (ggz 36.10c. 101 Table XIV. Temperature dependence of the exchange time in EDAl and corresponding relaxation rates in the absence of exchange.a T(°C) T(msec) l/T2A(sec) 1/TZB(sec) 19.9 6.65 (0.16)b 272 283 25.5 4.353 (0.08) 237 247 31.3 2.851 (0.03) 211 222 36.8 2.042 (0.02) 190 192 42.6 1.376 (0.009) 170 170 48.3 0.9508 (0.007) 153 151 53.9 0.6633 (0.009) 139 136 59.5 0.4856 (0.01) 127 123 64.9 0.3462 (0.003) 116 112 70.3 0.2500 (0.002) 106 102 a) 0.15 M C222 + 0.6 M NaBr. b) Standard deviation. 102 Table XV. Temperature dependence of the exchange time in EDA2 and corresponding relaxation rates in the absence of exchange.a T(°C) T(mseC) l/T2A(sec) l/T2B(sec) 30.4 2.338 (0.03)b 256 270 36.2 1.603 (0.02) 189 191 38.0 1.413 (0.01) 182 184 41.4 1.126 (0.008) 171 171 42.4 1.048 (0.008) 168 167 45.0 0.8913 (0.008) 160 159 50.2 0.6154 (0.004) 145 143 53.8 0.4862 (0.004) 136 133 60.8 0.3111 (0.003) 121 117 69.0 0.1918 (0.002) 106 101 76.4 0.1238 (0.001) 93.9 84.0 a) 0.30 M C222 + 0.6 M NaBr. b) Standard deviation. 103 Table XVI. Temperature dependence of the exchange time in EDA3 and corresponding relaxation rates in the absence of exchange.a T(°C) T(msec) l/T2A(sec) l/T2B(sec) 19.3 2.154 (0.04)b 271.5 288.3 24.4 1.448 (0.02) 243.8 255.0 30.6 0.956 (0.02) 215.1 221.0 36.1 0.600 (0.02) 193.3 195.8 41.9 0.486 (0.02) 173.4 173.4 47.4 0.3586 (0.005) 157.0 155.3 53.3 0.2295 (0 002) 141.6 138.7 58.8 0.1626 (0.0007) 129.1 125.4 64.4 0.1162 (0.0007) 117.9 113.7 70.2 0.08232 (0.0005) 107.6 103.2 a) 0.45 M C + 0.6 M NaBr. 222 b) Standard deviation. 104 ----~----3-~--fi---~5----S---‘5----§-'--S----5-'--5-"‘S-’--§----5-°-'$--"§‘---5"'-5'---§-°--5-"-R I: n x l 3' A fl .1 v a: 3n I J! J\—-n-n Man—fun..— l x I O I c h 5 l: I =0 I [=0 ammo o x xxx no I x u: n S “no ‘I I I x I o l 3 I no I - 1m 3 [I g I l I . a It! I 0 M! S :( an R I 0= I: l I I = ( :rrnx I I ("PA (11:: I l I I :q:t==l I g;0;fi:| 1117' I 1:::;: I ) nHOIIH‘:-;: :% I ------------- ' --------- ‘ ----- ' ----- \----§----5----9---—S----9----9----S----S—---S----S -------- S--~ofi| 00000000000000.0000. ( “)0 - )5 ----S----S----S----S--°‘<-~-~S----S-'--S----S----§--vo§----§----5--‘-5—---S----S----S----§-°~-§----S 0 l 1 L l 0 I N h h u I ‘I C 0 9 g g l '1 0 0 g I l I g 0 O X l 0 “ I h l 0 OX 1 O u I: ‘v I ooonnjnnqn.\'m:a:on:on u, I: lxxxxlxuuu H u-I'M') 1 1:1 Xxl==finh I =20 — IKXI==30 XJI‘IIIIT::0::H l:::::::l [X I ”onA-' 30 O: OO::::n :ITVT:: :‘ I'~--‘----S----9----1--‘-§—-—-$----<----<----9---~9---—S----§----9---'5-—--9-—--$----S----S—---<~¢--\I coco-00.000.000.000. r Figure 26. Computer fit of spectra obtained with 0. 2 M C andO 0. 4 M NaI in H20. (a) 23. 6 C; 222 (b) 3. 30 C. 105 I----S~---‘----§----S----$----S----So°~-9----§---o§o---S----§----fi----%----9~---§~---fi—-—oC-~-—<----%I I I I I lfi- I t 0 O l I l I ' S I [O I I I I n O I I I I ‘ 9 I 0 I I n I. I l I O I I S I U I I I I 0 I I I l I ‘ Q = I I v I 0 I a, s I I H 0 n l l I . . l ‘3 ‘I t I =’ I I I 3 O I ‘y 1 CI 2 r I D I = In I 0 II I ' 0‘ )l S I 0 0- I I n l - I I on In I I :1 I-n I Q 2: 1‘- l S I In: fn‘ll I I I n")! 00.0!!! I { Inn:n:t xwfinn ;II I l I [too - :‘l x 21))1 0 '1' - I 1| {uII I 0 na '_,__' .......... q---,r_-__v‘-_,,g-,__rl_--_r'_,<-’-__1<‘__- r--__r _____ r)__--r ,,,,, r _____ f _____ q- _-r__ """"I OIOOOOOt J)¢bIJO'I.)O f‘ 4;]. - )‘I _-_-S__--:.-___r-)-___-c‘--__r _____ (’-_--r,__-_<')____s)_-_-c‘___,rl_-__r.'__-~r)___-< .... r"____(‘__-,l‘__-_t _____ (‘ ___l' x I l I too I O = I ' 9 I I O I = I l I ‘; Ox Q ‘ I o I O I I 1, I t. n I I : o . I I I : a. I I n l 0 l: I b w ' I I In I : Inn I x (m I l "I, 1 H I'llfifl \ I s I I I l n I y “0 I I n I" I '5 l I IL, I n I I 0 ID I I I In I I ' In I I" ’, f) I : I I n - I l 0! t l | n: :n- J ‘. l ht Irm . I 1:!) If? I l I flan :nfilv ) I I .1 urn-II l' I 1 fin") ll find-Wnl I I I II 1 ‘01]! l II'III'II Iho'y- -'. ,_,_' _____ '--';----r.--—<'.--—-',—---'.—--—r)----S----'|'---'\---" _____ g_-__l)____t‘__-_l'____\_-__I____I.7--, 'I ‘0..00'I(.’)IHOD...OI ( ppm - 1"] Figure 27. Computer fit of spectra obtained with 0.2 M c2 2 and 0.4 m NaI in H o. (a) 39.3Oc; ? o 2 (b 26.9 c. 106 delay occurs and one relaxation time is much shorter than the other, as in this case, some of the intensity of the faster relaxing site will be lost. As a consequence, the observed intensity of the broad peak will be lower than that of the calculated. Spectra from the PYR solution were analyzed with Equation 5.3'. In the slow exchange limit, the A param— eter was dropped and MA and w along with the usual parameters B were allowed to be adjusted to obtain the best fit of 5.3' to the experimental data. Some fits for spectra from PYR solution are shown in Figures 28 and 29. Spectra from the THF solution were also fit with an unmodified form of 5.3'. A few of the fits for THF are shown in Figures 30 and 31. Variations of T with temperature for H20, PYR and THF solu— tion are listed in Tables XVII and XIX. 5.3 Results and Discussion 5.3.1. Some Sources of Systematic Error Many sources of systematic error occur when the Fourier transform NMR technique is used. These can cause distor- tion of the observed lineshapes. Some sources are pulse feed-through, the first—order phase correction and narrow— band audio filtering. The effects that they produce can sometimes be eliminated instrumentally and/or the calculated lineshape can be modified to include these effects. For example, the calculated NMR lineshapes included the first— order phase correction as described previously. 107 ----‘--~-S--c-S----S-o--$---oS.---$--nog-c---S--o-fi--o-Soo--\.o-.5--o0‘--o-5--oo)----‘----‘----‘-¢--§l l u I 0 I 0 Jr 0 O 10 z I o I O! l O I O I . - . o - l I t o l 0 0 I ‘ l 0 £ l l I K 0 “‘ " “ I ' I I ‘0 I I ° 1 X 61 n I x g - S O 0 I I I O l I J n x o I v Q n a a x X l 0 I x I J l y 0 0 0 0 “ I = O! x xx 0‘ 0 0 0 l 100 1 IX 0 ‘1 I: [(1:0 :0: = ‘ IQ x0=zn= l : l O 0 l0 I l: l ‘ [In tojl ‘ Aljn 10:0 ' I on l=0=010 l ‘00: X l 11‘ ::II:O I ‘0 ' ‘ I --—-s----s----s-—--<----<----s----a----<----s---—<----s----c----s----s----<----e---—S—---s---—s--——sl DC...OOOO\OOI)O0.0000 r )yu-;) l ' 0 z I ‘b 0 l X ‘ o I O X b “ I ' a I a O X ‘I I ” ' 1 I I l 0 0 I l l x x I i . . 0 l i S l n: O I CI 1 X 3x l- ‘0‘ 0: I l ') 10::nlII l l x::: I fla‘ lflnfl=3rh =0 ¢,::2::I:l -_-_‘, ---'\--- ‘,---—S————‘\-o——')—---',---—'\—o-—c|---—“—-—- CocooouoooooOo.0030 C VIII-"7 Figure 28. and 0.4 M Na¢4B in PYR. 66.7OC. 23%? (v—---§----S‘--‘5--' (a) :0 IO l-‘OO ll:’)00 lllf‘vzzn'NIKI qum 2 -9----<----s_---<----g---- Computer fit of spectra obtained with 0.2 M 93.20c; ----%----9--—-e--——s---—a----s----e----s-—-—s----s—---s----s----s----s----s—--—s—--—5----s--v~<~---sv .— : —-.——-J‘\—.——— fi———n-¢J§—n—n—-—o]¥<——n— _._-_H _ ___, ,9 ,0 _ _ _,________,. _ D _ t 3 x I ' ' ’ ' . '"‘ " "" 1 " OX 1 0 . . ‘= _ ..... _. __ _ _ ' ' ' Ill 2:0 I S l1:00 x—Oan IO— -—= 1!:fifln::(2000 . L I — ( :::x=g:nnnno ----- -:::fl:;::a;::-.{)n [[1 l XIIIA----— N l I .---~s--v-s----s----5----s----s----s----s I I I I I g I I l ‘v I l a I I I : : 0000.00.60.0000100 O r Figure 31. 110 ‘Xfi- 00’)! I 0::::::::O , . ::::::0=:=x x l 1----s----s---—s----s----5----s----5--——5----s----s--—-s----s-—--s----s----5----s----s----a----s IN" - S7 I----s-—--s----e----s-—--s----s----s----s----s----s-——-s----s---—s----§----s----s----5—---s----s----$ I l_ - _- ..__,_ - __.r._-,-..-,,ooo ==i _. _ , _ _ _ I x x no: 4 O I n I { -.-.___-_. - _--_-_____..-,-..- i.-.“ ___.-__-OL-_.- -__O._.__-_. _._ -- - _ ._ _ l l X l. _ rue i i __ ____-_._ -_-~_.__O ...... _ __ - _- i _ ‘ I I - f- -_ - -- ----- .__ 0.. _-...___.-_.-.- - _ _ - _ _- - ‘ X l _ _ _ ,___.---_.-____-_____._.. -.___.---__._._-O , _. i _ i. _ i__. I = - - .fi- _ .H..- _” -_ _r_ -- I _____ __ X 0 f. _____ *7 _ ________ _._ _ _ _ _ 0 _ - , _ __ _ X l O _ _. ._ - __ -_ 3 _ _ I, X I l l . ,, _ _ 0_ _ _ __fi _-, 0 _- l X c 0 n _ - _- ._.____ _A___ _____ _X 2 l' c X F A A___ --i _ -,-,___, __-__ _ _ ,rr_- _ 0 , _- 3 h 0 l 01 I‘ - =X ---._ i __. _- ___ __ O _- I. on 110:] O=Tflx l 1!)-0’:=::) ll l l::-00:-1::3 f ( .’---_ ,--.-- x - _ 0:0 I I =O::nl II I fl::;n:fhfiw: l ““1020?! x I r—--—n----t.--- ‘~---,---—s---—s—-——9--—-5~---9----s-——-s----S----s----S----h I 1H! (33 p00000|0900u¢oo¢0 0 (__' —-—.- ‘ I l ‘(1 l---- f -..—- ,---- Computer fit of spectra obtained with 0.2 M C 2 and O. 4 M Na¢4B in THF. (a) 56.8OC; (E)? 62. 80 c. In W.-- ---'- T-Lté----s----s_-‘.‘.§----?,----s---;5’::§.--.3---3;--.:g---.3::-.§-.'::§—.--:g‘:.;-5:-2- ::.;gl 0 0 ' -.- ~ W- ' -' -_—-_ x m ’ ll_‘-_— -.———“ H_ ---_“—. ___~-._ 1 0 10 1000 l 1 lo I ___._._ M___-___.-.__i_____ __ I!XO___ -_ _a._._ “"i_”. --_.__._ I J . ° ....... _ ..-_-__--__-____--_. __ ______,-‘._-____ __ ,-_-- -«q-.. r- i- -. 0 _-_ -____ _ -~___ __ R“_ _ _ ______ 1 0 —’ ' ” l O I O __._ A __ ___._. __ _. .___.-___ __ J __- ___________ ___-i-___ _ .— t I l _ -.. __ _ -i -_ __ ____._ ______._____. -3..,_._-,____-________,___ .__r ,_-._____,. I .. I I I I _._..___.__ p... _. _____-_.- i _- _ __-__ -__-______ ____ __‘._._______.,, __ iii ___-___ .. Z x - _..__~-____ ._..___ i,___,__.____. _=._ _. _ _ O -_,,.. - ~i_ _ _ - __-_._ S I I X I ____-_ - r __ ____ __ O.-_, -m ____ ___-,--__ .= . _ -_ _ .....___ __ _ l: a L X =0 3: = 00 l - _ - _-_ _ U 0: 3 _. __ .u- ____.____,___.-._ __ _~ 11:00 _ _ 111 Table XVII. Temperature dependence of the exchange time in H20 and corresponding values for relaxation rates in the absence of exchange.a T(°C) T(msec) 1/T2A(sec) 1/T2B(sec) 3.3 34.5 (3.1)b 61.0 503 12.0 11.34 (0.4) 49.4 389 22.2 4.335 (0.07) 42.3 294 23.6 3.830 (0.05) 41.5 284 26.9 2.877 (0.04) 39.8 261 29.2 2.178 (0.02) 38.7 247 30.6 2.030 (0.02) 38.0 238 33.8 1.616 (0.02) 36.8 221 37.0 1.120 (0.02) 35.8 205 39.3 0.892 (0.01) 35.4 195 45.9 0.5122 (0.009) 32.3 168 52.3 0.3333 (0.008) 31.3 147 a) 0.2 M C222 + 0.4 NaI. b) Standard deviation. 112 Table XVIII. Temperature dependence of the exchange time in PYR and corresponding relaxation rates in the absence of exchange.a T(°C) T(msec) l/T2A(sec) 1/T2B(sec) 55.2 32.0 (4.3)b 83.8 115 66.7 21.9 (2.1) 78.8 102 73.3 15.2 (1.4) 76.3 98.0 81.4 9.82 (0.5) 74.8 88.4 93.2 5.31 (0.2) 73.3 80.4 98.7 3.82 (0.1) 73.3 77.8 102.0 3.157 (0.09) 73.3 77.4 105.4 2.815 (0.06) 73.8 77.4 109.6 2.333 (0.06) 73.8 76.9 116.4 1.609 (0.02) 74.1 76.5 125.8 1.013 (0.011) 74.8 75.4 129.4 0.8607 (0.008) 75.8 75.4 135.4 0.6611 (0.006) 76.8 74.9 142.1 0.5051 (0.006) 77.8 a) 0.2 M c222 + 0.4 M Na¢4B b) Standard Deviation. 113 Table XIX. Temperature dependence of the exchange time in THF and corresponding relaxation rates in the absence of exchange.a T(°C) T(msec) l/T2A(sec) l/T2B(sec) 35.5 24.2 (1.2)b 108 146 39.4 20.2 (0.8) 106 139 44.1 15.0 (0.4) 105 132 45.8 12.7 (0.3) 104 129 49.4 10.2 (0.2) 103 124 52.2 8.43 (0.1) 102 120 56.8 6.089 (0.09) 100 114 59.1 5.334 (0.08) 99.3 111 62.8 4.063 (0.05) 98.1 107 65.0 3.510 (0.03) 97.5 105 68.2 2.876 (0.03) 96.5 102 71.3 2.363 (0.03) 95.6 98.7 74.6 1.872 (0.02) 94.6 95.7 a) 0.2 M c + 0.4 M Na¢4B. b) 222 Standard deviation. 114 It is not at all obvious by inspection of the spectra that a first-order phase correction is needed. This is be- cause a single-line spectrum can be phased to appear as an absorption mode signal by using only a zero—order phase cor— rection. Even in the case of a doublet, the spectrum can be visually phased to an apparent absorption mode signal without using a first-order phase correction (an example of this is shown by the spectra in Figure 17, in which the total change in phase from one end of the spectrum to the other is 44.5°). By contrast 13C NMR spectra usually exhibit many narrow lines which span the entire frequency range sampled. In this case Visual evaluation of the first—order phase cor— rection is easily made from a single spectrum. The first-order phase correction will have its maximum effect on the lineshape for an exchanging system at the slow exchange limit (low temperatures), where the spectral in— tensities are spread out over a broad frequency range. Figures 32a and 32b, which contain the same experimental data, show the effect of 8' upon the fit. The lineshape calculated in Figure 32a does not include a first-order phase correction term whereas that calculated in Figure 32b includes the proper first-order phase correction. The fit in Figure 32b is slightly better than in 32a. Failure to make such a correction usually only causes small differences between the calculated and observed line- shapes. However the effect is magnified for large chem— ical shift differences such as we observed with EDA solu- [-.--S--ou§~---S—--.S--cog-c.cS----§----<-~-—5--~-fl-o--5----§----S---os----S-n-n‘ooo-S-o-o§—-—-s----5 O n! l I ‘b n I I) l O I I c I '> I O a I, 1 t. 0 ‘t l " 0 l I 0 l l I ‘. l S l 0 I O 03 1 a o l:- I l I I ' x - OI 0‘ a 1 ° l u n: M I V I I OI 0! 1 1 -o m 1 , 2 130001000; 0 l S l : IIIIAII nnlI | l u 00!! l I x' 0'): III I I not). ~r - | <::::=::-: .:;;:::= -:—-_-;= :r‘ i—---‘)—----- -"---- -- —v--‘.----§----5----fi----§--ovS----S-—.-S----‘.----H] ----5----«----s---_S-- q .«ulm . .I‘M (9'0! 25.5 _ -_- --- I CHOOOOOC).QOOQIIOOOI C I----5-.--<,----q----¢,----s----§----c.----s----5-_..-q.---<----5----q---_5----q----s----5----I:,----3----¢, 0 3‘ OI I I I I l g I I I I g I; i 1 S l i I, l I I I g I 1 1 I tl I I I l I, I I I I c. i l l l I; l ‘ I I ‘ z I l t 0:): n! ' I, : t l I n t O! — 1 I l “ : n! ' x I r: 0 - x—--:: :I - n:— D: :::T= x::_- ==;-..—.-:::-:.—:nnn ‘9=='I====?=“—=:====') ‘11::.;; : ----‘1' --‘----,----'.--~— -‘--'>----‘."'-'\-'-~‘~"-‘5-*--‘---~-‘~----§----'~.--'—S----"-°--‘3----‘,""5----‘x ooooooonaooooooaoooo.o ( ‘A‘QLHA o .|','t Cowl 25.8 Figure 32. Computer analysis of the 23Na lineshape for a solution containing 0.15 M C 22 and 0.6 M NaBr in EDA at 25.5 C. (a) No first-order phase correction; (b) first-order phase corrected. 116 tions. In addition as shown in Figure 33, straight line Arrhenius plots result in either case, but give different activation energies. The experimental data are the same for both sets of calculations shown in Figure 33, but the T values were evaluated with and without first-order phase correction. This shows that failure to make the proper first—order phase correction might not be readily apparent but it can have a pronounced effect on the results. In this case (EDA l), neglect of 0’ correction gives an 8% error in the calculated activation energy. Another source of error can come from the 4-Pole Butter- worth active audio filters which are frequently used in Fourier- transform NMR spectroscopy. All the experiments discussed were performed with a sweep-width of 5000 Hz. To obtain an optimum signal-to-noise ratio, a narrow band filter (0 - 5000 Hz) was used. A perfect filter would be one which passes any frequency between 0 and 5000 Hz and has an infinitely sharp cutoff at 5000 Hz. However, no such filter exists. The attenuation characteristics of the filter employed in these experiments is shown in Figure 34. An input sine wave which is passed through the filter at 1 volt (peak to peak) and 5000 Hz frequency is attenuated by 40%. The attenuation at 3000 Hz is only 3.1%. The intensity function shown in Figure 34 can be matched to within better than 1% with an equation of the form I = ax2 + bx + c, 117 ‘Q\ i\ o 103» 5 ... ,r 1* U _. an U) U .x 2 10 ~ _ .1 I J l l l 1 ' l l 2.9 3.0 3.1 3.2 3.3 3.4 103/T (°K‘1) Figure 33. Arrhenius plot of k (rate of release of Na+ from C222) for EDAl solution. (0) No first-order phase correction; (O) first-order phase corrected. .umuaflm m>fluom nuuozumnusm maomus um ooom m How mocmsvmum MN 0H\H wuflmcmucfl omumscmuum mo uoam .vm muswflm ANIVme x .09.“— 118 3.. o.» o.~ 3 ed _ fl _ _ L ed .. n I/n. mwguml /. .. /./. .. /o/ ./In. 1. 1.. 3 119 where x represents frequency and I is the intensity. A least- squares fit of the above equation to the data in Figure 35 was performed to evaluate the coefficients, a, b and c. This calculated intensity function was then inserted into Equation 5.3' as an amplitude adjustment. The corrected intensity form of Equation 5.3' was fit to the EDA 3 data (these data were used since, of all the spectra, they had the most spectral intensity at the highest frequency and this is where most attenuation would occur). A comparison of T values calculated from the best fit of 5.3' with and without this intensity correction to the observed data from EDA 3 solution is given in Table XX. From these results it is clear that no intensity corrections are needed for this work. Another source of error comes from the digital sampling which is characteristic of a Fourier transform spectrum. Even if an ideal spectrometer with no baseline distortion from pulse-feedthrough were used, baseline artifacts could occur because of the digital sampling. If we are interested in a spectrum with frequencies as high as 6 Hz, we must sample the free induction decay at a rate which is at least 26 in order to properly represent the spectrum in digital form (this is known from information theory). If we sample at a rate 26, and frequency of 6 + e (where e is positive) is examined, then there will be less than 2 points per cycle representing the frequency 6 + e and the frequency 6 - s also has the same value at each of the sampled points. Thus the frequency 6 + e is indistinguishable from the frequency 6 — e , and any 120 Table xx, Comparison of exchange times with and without intensity correction for EDA3. Temp(°C) T(msec) Tcorr(msec) 19.3 2.154 (0.04)a 2.151 (0.04) 24.4 1.448 (0.02) 1.447 (0.02) 30.6 0.956 (0.02) 0.956 (0.02) 36.1 0.600 (0.02) 0.601 (0.02) 41.9 0.486 (0.02) 0.486 (0.02) 47.4 0.3586 (0.005) 0.3584 (0.005) 53.3 0.2295 (0.02) 0.2293 (0.002) 58.8 0.1626 (0.0008) 0.1625 (0.0008) 64.4 0.1162 (0.0007) 0.1161 (0.0007) 70.2 0.0823 (0.0005) 0.0823 (0.0005) a) Standard deviation. 121 information of the former is aliased to the lower frequency 6 - 8. Taking a simple case, if a spectrum is obtained which has a single peak which is broad (say a factor of 20 times smaller than the sweep width) then baseline distortion will occur and might be observable with a good signal-to-noise ratio. Any intensity on the wings outside the sweep width is folded back into the spectrum. The dispersion mode spectrum is much more distorted than the absorption mode spectrum because of higher intensity in the wings. This fold- ing back of the wings causes a baseline distortion. If a phase correction is performed, the larger distortion folded into the nearly pure dispersion mode spectrum is mixed into the phase corrected absorption mode spectrum. Foldover from the wings of the dispersion line becomes worse as the width of the line is increased compared to the width of the spec- trum. Normally, the linewidth of a peak is smaller than two percent of the total width of the spectrum and the signal-to— noise ratio is smaller than 100, so that the spectral dis- tortion is negligible. In conclusion, even a perfect spec- trometer can give Fourier transformed spectra which are base- line distorted because of digital sampling which leads to only a finite range of distinguishable frequencies. 5.3.2. Mechanism of Exchange Two possible mechanisms of exchange were considered. The first (I) is given by Equation 5.12 and assumes exchange 122 proceeds through an association—dissociation process. k 1 + (Na+) : (Na+C222) 5.12 k—1 (C222) The second mechanism (II), given by Equation 5.13, is bi- molecular and also represents the overall exchange process. * + + * + + (Na ) + (Na C222) 2: (Na C222) + (Na ) 5.13 Results of Lehn et 31.39 combined with those of Ceraso and Dye4O indicated a preference for mechanism I, but mechanism II was not shown to be inconsistent with the data. The com- plexation of crown polyethers with sodium ions proceeds via mechanism I. To determine whether mechanism I or II is con— sistent with the data at all temperatures, we examined the concentration dependence of T. The data from EDA l, 2 and 3 were used since PA varies as 0.25, 0.5 and 0.75 respectively. The relationship of T to a mechanism is made through the definition of Ti l/T * rate of removal of nuclei from sites of type i i number of nuclei in sites of type i 5.14 From Equations 5.9, 5.10 and 5.12—5.14 we obtain the following equation for T. Mechanism I T = PA/k—l 5.15 123 Mechanism II T = 1/2k2(Na+)total Mechanism I predicts a dependence of T upon the relative population of site A while mechanism II predicts a dependence of T upon the total concentration of sodium ion. Table XXI gives the calculated activation energies for each solution from the best fit of an Arrhenius expression to the data. To test for consistency with either mechanism I or II an exchange time, Tc was calculated from the average value alc’ of the activation energies and relaxation times of all three solutions. Then Tcalc values were calculated for equal popu- 1ations, PA = PB = 0.5, at each temperature for which the experimental exchange time, Tobs' had been evaluated. The ratio of Tobs/T is plotted against temperature for each calc solution in Figure 35. It is clear that the data at all temperatures are consistent with mechanism I, since the values are proportional to PA' Measurements with solution EDA 2 were performed before the importance of an instrumentally determined first-order phase correction had been realized and a computer estimate of 0' was used (previously described) to calculate lineshapes. This might be the cause of the systematic deviations shown for this solution. Activation energy plots, log k_l vs l/T, are shown in Figures 33, 36 and 37 for EDA l, 2 and 3 respectively. Ac- tivation energy plots for H O, PYR and THF are shown in 2 Figure 38. Activation energies, rate constants (k_l) and values of AH:, As: and AG: for the release of Na+ from the 124 O C O 1.5 0—o*—.~—-———.—o . . 0. .g. o ’ o O O .— .2 10 9 8 L > ‘8 I- O: C . .— U o o 0 . . ° ’ I l l l I l 20 30 40 5O 60 70 t(°C) Figure 35. Plot of the ratios TobS/T vs temperature (°C) 1 . for EDAl, EDA2 and EDA3 sgiugions. 125 103 __ . \ 1'1 0 k (seC°1) I I / 2 l 1 J 1 I ' 1° 2,9 3.1 3.3 103mm") Figure 36. Arrhenius plot of k (rate of release of Na+ from C222) for EDA2 solution. 126 .l (3 u k(sec4) '\ 2 1 I I 1 1 1° 3.0 3.2 3.4 103/T(°K") Figure 37. Arrhenius plot of k (rate of release of Na+ from C222) for EDA3 solution. 127 _ \ 0 H20 0 D THF 103,. \e 0 PYR B :'\ \ ._ R. 20 7 \ \3. ‘00 3 - ‘\ D. \ O. a“. 0. CE] 0\ x C D O 1 \ . O 10 )— . E I D 1— \ : \ 80b 0 ~ '\ \ 5 ‘1‘ \1 O l L, I l l l 2.4 2.6 2.8 3.0 3.2 3.4 3.6 10‘} (°K“) 'I Figure 38. Arrhenius plots of k (rate of release of Na+ from C222) for H20, THF and PYR solutions. 128 cryptate complex are given in Table XXI. Rate constants were calculated from T values by assuming that the pathway of sodium ion exchange for H20, PYR and THF solutions also proceeds via mechanism I. A complete error analysis, in— cluding the cross correlation terms, which account for the coupling of the parameters was performed in all cases. The free energy of activation is directly determined from k_l and has the smallest standard deviation. As expected, the average value of the activation energy for EDA solutions (listed in Table XXI) is in good agreement with the value of 12.2:l.1 kcal obtained by Ceraso and Dye4O 23Na NMR techniques. For the H20 solution a forward rate of association k1 is calculated from 6 1 -l by using continuous wave k_l and Ka to be 1.2 x 10 l was reported for sodium ion exchange in D20 solution at 314°C and is in fair agreement with our calculated value of k_l = 16 sec-l, at 3°C for sodium ion exchange in H tion. 20 solu- Activation energies vary by nearly 4 kilocalories, in- dicating the definite presence of a solvent dependence. A variation in activation energy with solvent for Li-cryptate 135 They decomplexation has been reported by Cahen et 31. observed a rough correlation between activation energy and donicity of the solvent as expressed by the Gutmann donor number (D.N.). As solvent donor number increases the activa- tion energy for release of Li+ from the cryptate complex increases. In our case, no correlation is indicated (the M- sec . A value of k_ = 27 sec- 1 129 .coaumw>mp pumtamum Am Amo.ov vv.va Am.ov m.n I Am.ov m.mH Am.vv o.mma Am.ov m.mH >¢I.ma 0mm ”No.0v NN.@H A©.ov H.@ I Am.ov m.mH Anm.0v mo.m Am.ov v.wa mmB Avoo.ov enm.ba Ao.ov m.mHI Am.ov m.ma Amo.ov ¢H.H mAm.ov N.va mwm “mommmv HImHoE HImHoE “Mommmv HImHoE ucm>aow HImHOE Hmox HIM oamo Hwox Hloom Hmox Koo< x m< x m< Hlx m .mucm>HOm msoHHm> ca mmcmnoxm wumummuo EUHUOm mo mnmumamumm oafimcmpofinmnu pom mmpMH mmcmnoxm .Hxx magma 130 donor numbers for THF, H O, PYR and EDA are 20.0, 33.0, 33.1 2 and 55.0 respectively). In both the lithium case and our case, the exchange rate in PYR solution is very much slower than the corresponding rates in H20 solution. By contrast, Shchori gt gt.136—138 found that the activation energy for release of a Na+ ion from certain crown ethers was independent of solvent (to within 11 kcal). They suggest that the energy barrier to exchange may be determined by the energy barrier for a conformational twist of a crown molecule. However, their data are restricted to three solvents with very similar donic- ities and more solvents are needed to test their hypothesis. The positive entropy of activation of H O is significant. 2 It indicates that solvent participates in the transition state. Values for entropies of transferAStr of univalent electrolytes from water to other solvents show that the standard partial molar entropy of the cation in water is higher than in other solvents143 (after corrections have been made for differences in dielectric constants). This is due to extensive structure-breaking of bulk water when an i in H20 is therefore a strong indication that the solvent participates ionic solution is formed. The positive value for AS in the transition state. VI. ALKALI METAL NMR STUDIES OF SODIUM, RUBIDIUM AND CESIUM ANIONS 6.1 Introduction It had been evident for some time that solutions of the alkali metals in amines and ethers contain species of stoichi— ometry M- (see Chapter II). The shift of the optical absorp— tion band with metal, solvent and temperaturezg'30 suggests strongly that the species is a centro-symmetric anion. How— ever, other models cannot be ruled out on the basis of optical evidence alone. Figure 39 shows three other contenders for the M- structure. Indeed, one or more of such structures might be responsible for the diamagnetic species in metal— ammonia solutions which, to date, show no specific evidence for the existence of centro-symmetric anions. 6.2 Magnetic Shielding Constants of Alkali Metal Ions 103,104 102 99-101 Optical pumping, atomic beam and NMR techniques have established the shielding constants of the 133CS+ aqueous cations 23Na+, 87Rb+ and relative to the gaseous atom with an accuracy of at least 2%. Reliable 105 calculations of the shielding constants of Na+(g) and Na"(g) relative to the free atom have been made by using Lamb's complete expression for atomic diamagnetic shielding144 (first term in Equation 3.2) and analytic Hartree—Fock 145 wavefunctions. The changes in shielding constants are 131 132 H // \N/ \\ \N/ x/ \\ [I H\ :1 /H \‘ H\ :4 /,H = \ IH—N ...... M ...... Nqu H—N ...... M ...... NTH 92 IL...— H/ H l’ H/ i \H / \ ; \ / \ / \ /N /N\ \ \ /// \\ H I H H I H \e—a’ \ \1e5,,/ H EXPANDED ORBITAL ION PAIR WITH DIELECTRON MIG?) Figure 39. Three possible models for a species of stoichiom- etry M" (other than an alkali anion). All of these models permit solvent interaction with the outer p electrons of the cation. Ammonia is used to represent any amine or ether solvent. 133 relatively small, amounting to only 7.7 ppm (diamagnetic) for the addition of two 3g_electrons to gaseous Na+ to form Na'(g). Corresponding shifts for Rb_(g) and Cs‘(g) are also expected to be small. Diamagnetic contributions from solva- 109,110 tion are generally only a few ppm and the major effect of the solvent, both from theoretical expectations and experi- 98 mental results is a substantial paramagnetic shift (45 to 75 ppm for Na+) upon solvation. The magnitude of this shift correlates well with the donicity of the solvent,llllllzvl46';47 that is, the ability of the solvent to donate electron den- sity to the cation. It is via the interaction of this sol— vent electron density with the outer p orbitals of the alkali metal cation which gives the observed paramagnetic effect. 6.3 Results The key feature which permits us to study the NMR spectra of alkali metal anions is the complexation of the 32 cation by macrocyclic polyethers of the crown and cryptand 34 classes. There are two reasons for the importance of this complexation. First, the pronounced enhancement of metal 30'31'33'36 permits the use of metal concentrations solubility which are high enough to study by alkali metal NMR techniques. Second, the complexed cation is released slowly enough by the complexing agent so that, at low enough temperatures, the exchange of M+C with solvated cations is slow on the 222 NMR time scale as shown by Chapter V. This makes it possible 134 to observe separate resonances for M+C222 and M‘. Figure 40 shows a typical spectrum of Na+C222 ' Na- in ethylamine. On the basis of the positions and linewidths of NMR spectra of salt solutions which contain the cryptated cation, NaC+, the broad peak in Figure 40 is attributed to this Species. The narrow peak at high fields is then assigned to the sodium anion. Although detectable amounts of free Na+ cannot be present (because of reaction with Na- to precipitate sodium metal) its normal position in this solvent is also indicated on the figure. As expected from the solution stoichiometry the areas under the two peaks are equal within the error of their determination. The existence of two peaks rather than a single averaged peak, which would result from rapid exchange between NaC+ and Na", is expected on the basis of the results from the previous chapter. The 23Na NMR spectrum of Na+C222 - Na— has been studied as a function of temperature in three solvents, tetrahydro— furan (THF), ethylamine (EA) and methylamine (MA). The 87Rb NMR spectrum of Rb+C222 ' Rb- in BA and THF and the 133Cs NMR spectrum of 133Cs+C222 ° Cs“ in THF have also been studied. The results are summarized in Table XXII. 6.4 Discussion The spectra for Na+C222 ° Na_ in THF, EA and MA are compared in Figure 41. The most striking feature is the absence of a solvent-induced paramagnetic shift for Na- and 135 .OHumcmmEMHp mum mumenm m>fluflmom “Humz mdomsqm woumusuMm we mocmummmm .Uov.H pm A: N.ouv «m ca Imz . NNNU+mz mo :OHusHom m m0 Ednuowmm mzz MZMN .ov ousmflm Candy :25 .mo_Eo£0 Om. ow am 0v 0% 0m 0w 0 9.! Gm! _ m _ _ _ _ e d m _ . I E€§.§$I .fi {aékflfég + 0 02 + m2 IBI9U BI\ Mgsuoiul ImZ 136 He +.H+ as me- 62 .o+62 me.o -62 moe -- Aoe6ov 6.N+ -- -- 666 -62 em 4.0m- 229 6- 62 .o+62 mH.o 0+62 one-ome m.om- em H+ on se- 62 .o+62 me.o o+62 6.0m 6.66- as me- 62 .o+62 mH.o 6o+62 o.mm a.mm- 22s mm 64662 N.o +62 a.se 4.4s- «2 mm H62 mm.o +62 o.a m.ms- 42 me- H62 m.o +62 woe 0.6 N.H6- 0N2 mm e662 .66m +62 6m -- 626e6oveme- 0N2 mm .eees -um age .oee -- 6N666e- 0N2 mm .eees -2 «we Noe No.6 m.m6m.eem- o~2 mm .eees +60 Noe -- N.266.Hem- 0N2 mm .He68 +nm ewe Noe 6H.m enm.oo- 0mm mm .2268 +62 x>< 6 262V 2522c ucm>eom Aoov 22V 20H o.mom o.mmm Qx>< Am: m> >aom mama aoHumuucmocou .mcuow3m2HH one mucmumcoo mcfipamflnm wo umHH pmuomamm . H HXX OHQMB 137 0H 0.06- 229 as- -60 .o+6o e.0 -6o 00 0.000- 20020 mm -H .o+6o 00.0 o+6o 0v 0.6Hm- 20026 60 H60 60.0 +66 00H 0v 0.060- 002 mm H6o H.0 +60 me 6.6H- 229 06- -nm .o+nm H.0 :22 0mm m.0m- «m 00- -nm .u+nm H.0 -62 0000 006-2 mommo mm -H .o+62 0.0 o+nm 00me N0N- 0N2 mm -H .o+sm 0.0 o+sm 00m 0.00H- 20020 mm H66 0.0 +62 00H 06H 0.0HN- 0mm mm Hem H.0 +26 0v 0.0+ 629 6- -62 .o+62 me.0 -62 0-0 0.H+ <6 e+ 0» 0H- -62 .o+62 me.0 -62 600 0 2620 25620 666>Ho6 2600 220 60H o.wmm o.mmm Qm2< 62m: m> >HOmzvo dame coapmuucmocoo 26.22060 .HHxx men6e 138 .ccmummuo m.m.m ou muemeu U Ae .cofice mdoemem ecu we eueum eoceeemem Ac .cuox ucemeum ecu on Hemeu euec .peuflo we eoceuemeu oc euecz A0 .620262 ee62 66 suees edge u m>0 in .Eoum mdoemem Heucsec ecu MOM oflueu oauecmeEoumm ecu we > Use caewm oauecmes oaueum ecu 6H om eHecB .om Ablavr u 3 mc cecflmep .6 .uceumcoo mcflpaeflcm fie 26.uaoov .HHxx een6e 139 62 / / / / / J 1,” £th2 “113%?” 1153/712'41’8’4313’14”{I‘I’I'MF / THF 111 11111-1611111-11111 // / MeNH2 '5“va - "r‘ 4W “W“ 1 I _ l I l J l --20 O 20 40 60 80 100 Figure 41. 23Na NMR spectra of Na+C222 ° Na- solutions (=0.1 M) in three solvents. All chemical shifts are referenced to aqueous Na+ at infinite dilution. 140 the narrowness of this line. The complexed cation, Na+C222 has a chemical shift which is also nearly independent of solvent and is at the same position as for ordinary salts of Na+C222 in these solvents. By contrast the solvated cation, Na+, is strongly solvent—dependent (Table XXII). The chemical shift of Na- is not only independent of solvent, but is also nearly the same as that of Na- in the gas phase. This is completely different from that of any ion with filled outer p orbitals (including halide anions) as indicated in Table XXII. As discussed in Chapter IV, a paramagnetic shift of an ion is caused by the overlap (KY model) of outer s and p orbitals of the solvent molecules with the outer p orbital of the ion. Inner orbitals are not considered, since they are tightly bound and do not extend appreciably between nearest neighbors. The absence of a chemical shift for Na“ shows that the 22 orbitals are well shielded from the solvent by the presence of the filled 3g_orbital. This would not be the case for any model shown in Figure 39. We conclude therefore that the most reasonable model for Na‘ is that of a centro-symmetric anion with two electrons in the outer s orbital. It is not obvious why the shift is as small as it is, since interaction with the solvent to cause mixing of s and p character could yield a paramagnetic shift. Perhaps the large size of the anion causes the chemical shift to be small. 141 The extreme narrowness of the line for Na“ also attests to the high spherical symmetry of this species. Quadrupolar broadening of 23Na NMR lines is common (see Table XXII) and results from an electric field gradient at the nucleus. Even such presumably symmetric cations as Na+(aq) are quadrupole broadened to 5 Hz (full width at half-height). By contrast, the true linewidth of Na" in THF is less than 3 Hz compared with 23 Hz for Na+ in THF. Since the lines may also be broadened by the presence of e; via a paramagnetic inter- olv action, it is difficult to obtain the true quadrupole-broadened linewidth. The relative concentration of esolv tends to vary from one sample to another depending upon the solvent used and the method of preparation. Therefore, we find that the linewidths are not completely reproducible. Since exchange of sodium between Na‘ and either Na+C or Na+ would also broaden the line we cannot completely rule this out. However, it seems unlikely since the linewidth of Na“ does not decrease markedly as the temperature is lowered. Just as for Na‘, Rb_ and Cs- give narrow lines which are diamagnetically shifted from the corresponding cations by a large amount. The NMR absorptions of Rb+C and Cs+C were not observed in the M+C, M' samples, probably because of line— broadening. The signal-to-noise ratio was satisfactory for observation of the narrow Rb” and Cs‘ lines but not for the broad lines expected for Rb+C and Cs+C. Studies with salts indicate that the linewidth of Rb+C is far too broad to have 142 been observed in the metal solution case. The situation for the case of Cs+C is not as clear. If the linewidth were in excess of 200 Hz the signal would have been lost in the noise. In methanol at 250C the linewidth of Cs+C (iodide salt) is 30 Hz. It might be expected to be much broader in THF at —7lOC. The aqueous Rb+ and Cs+ ions are paramagnetically shifted 212 and 344 ppm from the respective gaseous atoms. Values for the gaseous anions are not known but are presumably shifted diamagnetically a few ppm from the atoms. The reson— ance positions of Rb’ in EA and THF are shifted 26 and 14 ppm paramagnetically from the atom and Cs“ in THF is shifted 52 ppm. Although the shielding constants are not as close to the gaseous anions as for the case of sodium, they are very close compared with the range of chemical shifts for the correspond— ing solvated cations.]-l0'148 The linewidth of Rb" in THF, 15 Hz, is much narrower than that of the Rb+ ion in any solvent. However, this is not the case for Rb" in EA or Cs‘ in THF which have linewidths comparable to those of the sol— vated cations. It is likely that the lines of Rb" and Cs- r b ad- solv or a e ro are either paramagnetically broadened by e ened by exchange. In conclusion, the NMR studies described in this chapter and the recent isolation of a crystalline salt of the sodium 36'37 are the most convincing evidence that genuine anion alkali anions exist in solution as well as in the crystalline state. APPENDICES APPENDIX A MODIFICATION OF RELAX 2 The data output portion of RELAX 2, a program written by David Wright128 of Michigan State University for computer controlled timing and data acquisition for two—pulse experi— ments was slightly modified in order to dump data from the Nicolet 1083 computer. The modifications are 143 Core Location Instruction 1310 2001607 1311 2001607 1312 2110577 1313 2404606 1314 2332600 1315 0005144 1316 0000001 1317 2000206 1320 0001346 1321 3110606 1322 3001326 1323 0001327 1324 2000206 1325 0001346 1326 0005051 1327 2124606 1330 1331 1332 1333 1334 1335 1336 1337 1340 1341 1342 1343 1344 1345 1346 1347 1350 1351 1352 1353 1354 1355 144 2000206 0001346 3110606 3001326 0001340 2000206 0001346 3001563 2001607 2134606 0006454 0162000 0000155 0001314 2404040 0000000 0000037 0001646 2165350 2001404 2001413 2125560 APPENDIX B PROGRAM CONVERT Program CONVERT converts data from octal to decimal in a form compatible with KINFIT. 100 10 105 110 15 115 116 999 PR¢GRAM C¢NVERT(INPUT=65,¢UTPUT=65,PUNCH=65) DIMENSI¢N IX(2), IY(2),X(2),Y(2) DATA(MASK=77777777777773777777B) READ 100,A,B IF (E¢F(5LINPUT))999,10 F¢RMAT (2310.2) READ 105,(IX(I),IY(I),I=1,2) IF (E¢F (5LINPUT))5,8 F¢RMAT (2¢10) PRINT 110,(IX(I),IY(I),I=1,2) F¢RMAT (1x,2310) D¢ 15 K=1,2 IF(IY(K).GE.2000000B)IY(K)=IY(K)+MASK X(K)=IX(K) Y(K)=IY(K) PRINT 115,(X(K),A,Y(K),B,K=1,2) F¢RMAT (2(20X,F10.0,F10.8,F10.0,F10.0)) PUNCH 116,(X(K),A,Y(K),B,K=1,2) F RMAT (2)F10.0,F10.8,F10.0,F10.0)) G¢ T¢ 10 C¢NTINUE END 145 APPENDIX C SOLUTION TO MODIFIED BLOCH EQUATIONS The Bloch equations which describe the motion of the X and Y components of magnetization in the rotating frame, when modified to include chemical exchange, are given by 1 G 1 dGA/dt + a G = -1YH + B-TA GA C1 A A 1MOA TB 1 _ . - 1 dGB/dt + aBGB _ 1YH1MOB + TA G A-TB GB C2 The solution to these equations for slow passage conditions is obtained by setting dGA/dt = dGB/dt = 0 and solving C1 and C2 for the total complex moment G = G + GB. Solving A for GA in C1 gives 1 (a + -—) A TA Substitute C3 into C2 and solve for G aBGB = -1YH1MOB + - TB GB C4 (a +-JL)T A TA A iYHlMOA (”lYHlMOB - WT)TB(GATA + 1) GB = C5 [(aBIB + l)(aAIA + 1)-1] 146 147 and from inspection in M . 1 OB ( 1YH1MOA " WHAWBTB + 1) B B [(dATA + 1)(aBIB + 1) -1] G = G + G = u.+ iv C7 inlMOB IMOA — GBTB+1)TA(aBTB G = {(-in + 1) lYHIMOA IMOB ’cfigfi;fif"°TB(aATA + 1)}/ + (—in {(aATA + l)(C1BTB + 1) — 1} ca T T . __ A __ B and for low values of the radio frequency field P M = M A O OA' PBMO = MOB PB G = ~inlMO{(PA + 3;?31T)TA(aBIB + 1) P + (P +—-—L—)T (0LT +1)}/ B GATA+1 B A A {(aAIA + l)(aBTB + 1)—l} C9 148 or G = -in1MO{(PAaBTB + PA + PB)TA + (PBaATA + PB + PA)TB}/ {(aAIA + 1)(aBIB + 1) — 1)} C10 Since PA + PB = 1 G = -inlMO{(PAaBTBTA + TA + PBQATATB + TB)}/ {(aAIA + 1)(0LBTB + 1) - 1} C11 or G = -inlMO{(TA + TB) + IAIB(aAPB + aBPA)}/ {(aAIA + 1)(0LBIB + 1) - 1} C12 as first obtained by Gutowsky, McCall, and Schlicter.139 To obtain the absorption and dispersion mode shape functions, one must separate the real and imaginary parts of Equation C12. TATB TA+TB Define T = G IA+TB+TATB(aAPB+aBPA) . = C13 —inlMO aAaBTAIB+1+aBIB+aAIA-1 Let Y = dividing numerator and denominator of C13 by (TA + TB) T T T T A B A B TA+TB + TA+TB + IA+TB(QAPB+QBPA) Y = C14 TATB a TA+TB A B 149 notePA+pB=1= f + +3 TA TB TA TB Thus 1 + T(a P + a P ) Y = A B B A (:15 TaAaB + PBaB + PAaA 1 1 Let k = ——-, k = ——— A TZA B TZB then 0A = kA-1(wA-w), aAPA = kAPA-iPA(wA-w) dB = kB-1(wB—w), aBPB = kBPB—IPB(wB-w) Let kAPA - KA, kBPB - KB and QA = PA(wA-w), QB = PB(wB-w) aAaBPAPB (KA—iflA)(KB-108) Then aAdB = P P = P P A B A B and C15 becomes 1+I{(k p +k p )-i[P (w -w)+P (w -m)]} Y: ABBA BA AB C16 (K -i0 )(K —i0 ) T[ A 2 P B B ] + (KA+KB)-i[QA+QB] A B 1+I{(kAPB+kBPA)-1[PB(wA-w)+PA(wB-w)J} Y = C17 (K K -0 Q -i(0 K +0 K ) A B A B A B B A . T[ PAPB J + KA+KB-1[QA+QB] 150 or Y = l+I{(kAPB+kBPA)-1[PB(wA—w)+PA(wB-w)]} (K K -0 Q ) (0 K +0 K ) A B A B . A B B A K +K +1 -1[(Q +9 )+T A B PAPB A B PAPB (K K -0 Q ) B P P A B R K +9 K T = 9A + QB + T( A B B A) P P A B U = 1 + T(kAPB + kBPA) V = T[PB(wA-w) + PA(wB-w)] Substituting C19-C22 into C18 gives U~iV (U-iV)(S+iT) Y=-——-.—-= S-lT 82+T2 Y (SU+TV) + i(UT—SV) ‘ 2 2 S +T and — _- - [-i(SU+TV) + (UT-SV)] G — inlMOY - +YHlMO 2 2 S +T C18 C19 C20 C21 C22 C23 C24 151 Define I = + Léfligyl' absorption mode shape function 2 2 S +T R = igglgyl, dispersion mode shape function 2 2 S +T and G = u + iv = yHlMO[R-iI] C26 where P P PP _ A B 1: [AB S——+—-—+ —PP(w-w)(w-w)] C27 T2A T2B PAPB T2AT2B A B A B P P A B T S = ———--+-——— + ——~———--'r0n -w)(w -w) C28 T2A T2B TZATZB A B P P U = 1 + T(TIL + TEL) C29 2A 2B P P P P B A A B T=P(w-w)+P(w-w)+ T[ (w-w)+-—(w-w)] C30 AA B B PAPB T2B A T2A B wA—w wB-w T = P w -P w+P w -P w + T( + ) C31 AA A BB B T2B T2A wA-w wB-w T=Pw +Pw—w+( + )T C32 AA BB T2B T2A and V = T[PB(wA—w) + PA(wB-w)] C33 152 V = TEPBwA + PAwB-w] In summary, the complex moment, G(w), is given as G(w) = u + 1V = yHlMO[R—11] R = 12%:§%L, dispersion mode S +T I = 1§§i§¥1l absorption mode S +T where P P A B T S=—+-———+———-—-I(w -w)(w -w) T2A T2B TZATZB A B P P U=1+T(-,i,—§—+-T-B—) 2A 2B w —w w —w A B T = P w + P w - w + T( + ) AA BB T2B T2A V = TEPBwA + PAwB-w] C34 C35 C36 C37 C38 C39 C40 C41 BIBLIOGRAPHY 10. 11. 12. 13. 14. 15. 16. 17. 18. BIBLIOGRAPHY W. Weyl, Ann. Phys., 191, 601 (1863). C. A. Kraus, J. Franklin Inst., 21%, 537 (1931). R. A. Ogg, Jr., J. Chem. Phys., 14, 295 and 1141 (1946). W. Bingel, Ann. Physik., 1%, 57 (1953). M. F. Deigen, Trudy Inst. Fiz. Akad. Nauk., Ukr., S.S.R., 5, 119 (1954); Jh. Eksp. Theor. Fiz., 26, 300 (1954). E. Becker, R. H. Lindquist and B. J. Alder, J. Chem. Phys. 25, 971 (1956). E. Arnold and A. Patterson, Jr., J. Chem. Phys., C011 3089 and 3098 (1964); Metal-AmmonIa Solutions, oque Weyl I, p. 160, G. Lepoutre and M. J. Sienko, eds. W. A. Benjamin; New York (1964). M. Gold, W. L. Jolly and K. S. Pitzer, J. Am. Chem. Soc. 84, 2264 (1962). Metal-Ammonia Solutions, Colloque Weyl I, p. 144, G. Lepoutre and M. J. Sienko, eds. W. A. Benjamin; New York (1964). 8. Golden, C. Guttman and T. R. Tuttle, Jr., J. B26 6Chem. Soc., 81, 135 (1965); J. Chem. Phys., 44, 3791 (16 C. A. Kraus, J. Am. Chem. Soc., 43, 749 (1921). R. R. Dewald and J. H. Roberts, J. Phy . Chem., 72, 4224 (1968). -—- mm E. Huster, Ann. Physik, 3%, 477 (1938). S. Freed and N. Sugarman, J. Chem. Phys., 11, 354 (1943). C. A. Hutchison and R. C. Pastor, Rev. Mod. Phys., 25, 285 (1953); J. Chem. Phys., éév 7959 (1953). “m R. C. Douthit and J. L. Dye, J. Am. Chem. Soc., 8%, 4472 (1960). _ ’— M. Gold and W. L. Jolly, Inorg. Chem., 1, 818 (1962). D. F. Burow and J. J. Lagowski, Advances in Chemistry Series No. J9, 125 (1965). 153 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 154 C. A. Hutchison, Jr. and D. E. O'Reilly, J. Chem. Phys., 3%, 1279 (1961). V. L. Pollak, J. Chem. Phys. 34, 864 (1961). D. E. O'Reilly, J. Chem. Phys., 35, 1856 (1961). s. R. Gunn and L. J. Green, J. Chem. Phys., g9, 363 (1962); J. Am. Chem. Soc., 85, 358 (1963). S. R. Gunn, J. Chem. Phys., 41, 1174 (1967). J. L. Dye, Metal-Ammonia Solutions, Colloque Weyl II, p. 1. J. J. Lagowski and M. J. Sienko, eds. Butterworths; London (1970). K. D. VOS, J. L. Dye, J. Chem. Phys., 38, 2033 (1963). L. R. Dalton, J. D. Rynbrandt, E. M. Hansen, J. L. Dye; J. Chem. Phys., 4%, 3969 (1966). K. Bar—Eli, T. R. Tuttle, Jr., J. Chem. Phys., 40, 2508 (1964). R. Catterall, I. Hurley, M. C. R. Symons, J. Chem. Soc., Iazz: 139- S. Matalon, S. Golden, M. Ottolenghi, J. Phys. Chem., 7%: 3098 (1969). M. T. Lok, F. J. Tehan, J. L. Dye, J. Phys. Chem., 16, 2975 (1972). J. L. Dye, M. G. DeBacker, V. A. Nicely, J. Am. Chem. Soc. 9%, 5226 (1970). “‘ C. J. Pedersen, J. Amer. Chem. Soc., 82, 7017 (1967). J. L. Dye, M. T. Lok, F. J. Tehan, R. B. Coolen, N. Papadakis, J. M. Ceraso, M. DeBacker, Ber.Bunsen-Gesellsch. Phys. Chem., 35, 659 (1971). B. Dietrich, J. M. Lehn and J. P. Sauvage, Tetrahedron Lett. 1269, 2885, 2889. J. M. Ceraso and J. L. Dye, J. Chem. Phys., 61, 1585 (1974). J. L. Dye, J. M. Ceraso, M. T. Lok, B. L. Barnett and F. J. Tehan, J. Amer. Chem. Soc., 26, 608 (1974). F. J. Tehan, B. L. Barnett and J. L. Dye, J. Amer. Chem. Soc., 26, 7203 (1974). 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 155 J. L. Dye, C. W. Andrews and J. M. Ceraso, Submitted to J. Phys. Chem. J. M. Lehn, J. P. Sauvage and B. Dietrich, J, Amer. Chem. Soc., 2%, 2916 (1970). J. M. Ceraso and J. L. Dye, J, Amer. Chem. Soc., 95, 4432 (1973). G. Lepoutre and M. J. Sienko (Editors), Metal—Ammonia Solutions, Colloque Weyl. W. A. Benjamin, Inc. (1964). J. J. Lagowski and M. J. Sienko (Editors), Metal- Ammonia Solutions, Colloque Weyl II. Int. Union Pure and Appl. Chem., Butterworth and Company, London (1970). J. Jortner and N. R. Kestner (Editors), Electrons in Fluids. Springer-Verlag, Berlin (1973). The Nature of Metal-Ammonia Solutions, To Be Published, J, Phys. Chem., 39 (1975). J. L. Dye, R. F. Sankuer and G. E. Smith, J, Am, Chem. Soc., 8%, 4797 (1960). H. M. McConnell and C. H. Holm, J, Chem. Phys., 26, 1517 (1957). m J. Acrivos and K. S. Pitzer, J. Phys. Chem., 66, 1693 (1962). m D. E. O'Reilly, J. Chem. Phys., 4%, 3729 (1964). W. A. Seddon, J. W. Fletcher, J. Jevcak and F. C. SOpChyshyn, Can. J. Chem., 5%, 3653 (1973). J. W. Fletcher and W. A. Seddon, Submitted to J. Phys. Chem. A. DeMortier, M. DeBacker and G. Lepoutre, J, Chim Phys., 380 (1972). M. T. Lok, Ph.D. Thesis, Michigan State University (1973). F. J. Tehan, PhD. Thesis, Michigan State University (1973). J. Kaplan and C. Kittel, J, Chem. Phys., 21, 1429 (1953). 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. R. Catterall and M. C. R. Symons, J. Chem. Soc., 13 (1966). R. A. Ogg, Phys. Rev., 6%, 669 (1946); J, Amer. Chem. Soc., 68, 155 (1946). T. L. H111, J. Chem. Phy ., 16, 394 (1948). R. H. Land, J. Chem. Phy ., 46, 4496 (1967). D. E. O'Reilly, ibid., 55, 474 (1971). K. Fueki, ibid., 50, 5381 (1969). K. Fueki and S. Noda, Metal-Ammonia Solutions, Colloque Weyl II, J. J. LagowskiIand M. J. Sienko (Editors), p. 19. Butterworths, London (1970). K. Fueki, D. F. Feng and L. Kevan, J. Amer. Chem. Soc., 95, 1398 (1973). J. W. Fletcher, W. A. Seddon, J. Jevcok and F. C. Sopchyshyn, Chem. Phys. Lett., 18, 592 (1973). J. W. Fletcher, W. A. Seddon, J. Jevcok and F. C. Sopchyshyn, Can. J. Chem., 51, 2975 (1973). F. Y. Jon and L. M. Dorfman, J, Chem. Phys., 58, 4715 (1973). R. Bockrath and L. M. Dorfman, J. Phys. Chem., 11, 1002 (1973). L. M. Dorfman, F. Y. Jon and R. Wageman, Ber. Bunsenges Phys. Chem., 75, 681 (1971). J. L. Dye, M. G. DeBacker and L. M. Dorfman, J, Chem. Phy ., 52, 6251 (1970). I. Hurley, T. R. Tuttle, Jr. and 8. Golden, J. Chem. Phys., 48, 2918 (1968). J. L. Dye, Electrons in Fluids, Colloque Weyl III, p. 77. J. Jortner and N. R. Kestner (Editors), Springer-Verlag, Berlin (1973). R. R. Dewald and J. L. Dye, J, Phy . Chem., 68, 128 (1964). D. S. Berns, E. C. Evers and P. W. Frank, Jr., J, Am, Chem. Soc., 79, 5118 (1957). 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 157 R. R. Dewald and K. W. Browall, J, Phys. Chem., 74, 129 (1970). K. Bar-Eli and T. R. Tuttle, Jr., Bull. Am, Phys. Soc., 8, 352 (1963). L. R. Dalton, M. S. Thesis, Michigan State University (1966). J. Eloranta and H. Linschitz, J, Chem. Phys., 38, 2214 (1963). ““ M. Ottolenghi, K. Bar—Eli and H. Linschitz, J, Chem. Phys., 43, 206 (1965). A. Gaathon and M. Ottolenghi, Israel J, Chem., 75, 286 (1971). ““ D. Huppert and K. Bar-Eli, J, Phys. Chem., 74, 3285 (1970). ”“ S. H. Glarum and J. H. Marshall, J, Chem. Phys., 52, 6251 (1970). “A J. G. Kloosterboer, L. J. Giling, R. P. H. Rettschnick and J. D. W. Van Voorst, Chem. Phys. Lett., 8, 457, 462 (1971). "‘—‘ ““' "“‘ m J. W. Fletcher and W. A. Seddon, Proceedings of Colloque Weyl IV, J. Phys. Chem., To Be Published. For references to experimental evidence see 8. Golden, C. Guttman and T. R. Tuttle, Jr., J, Chem. Phys., 44, 3791 (1966). M. Smith and M. C. R. Symons, Trans. Faraday Soc., 54, 338, 346 (1958). G. Stein and A. Treinen, Trans. Faraday Soc., 55, 1086, 1091 (1959). ‘“““‘ ““ I. Burak and A. Treinen, Trans. Faraday_Soc., 59, 1490 (1963). "”'" ““ J. L. Dye, M. G. DeBacker, J. A. Eyre and L. M. Dorfman, J, Phys. Chem., 76, 839 (1972). M. G. DeBacker, Ph.D. Thesis, Michigan State University (1970). M. G. DeBacker, J. L. Dye, J, Phys. Chem., 15, 3092 (1971). 158 90. T. R. Tuttle, Jr., Chem. Phys. Lett., 20, 371 (1973). mm 91. R. V. Pound, Phys. Rev., 1%, 685 (1950). 92. H. S. Gutowsky and B. R. McGarvey, J. Chem. Phys., 21, 1423 (1953). 93. N. F. Ramsey, Phys. Rev., 6, 699 (1950). 94. K. Yosidu and T. Moriya, J, Phys. Soc. Japan, 11, 33 (1956). m” 95. J. Kondo and J. Yamashita, J, Phys. Chem. Solids, JR, 245 (1959). 96. D. Ikenberry and T. P. Das, Phys. Rev., 138, A822 (1965). mmm 97. D. Ikenberry and T. P. Das, J, Chem. Phys., 43, 2199 (1965). ”m 98. C. Deverell, Progr. Nucl. Magn. Reson. Spectrosc., 278 (1969). 4, ’b 99. O. Lutz, Physics Letters, 25A, 440 (1967). mmm 100. O. Lutz, A. Naturforschung, 23a, 1202 (1968). _ ’b’b’b 101. O. Lutz and A. Schwenk, Physics Letters, 24A, 122 (1967). mmm 102. A. Beckmann, K. D. Boklen and D. Elke, J, physik, 333, 173 (1974). 103. C. W. White, W. M. Hughes, G. S. Hayne and H. G. Robinson, Phys. Rev., 174, 23 (1968). AAA) 104. G. S. Hayne, C. W. White, W. M. Hughes and H. G. Robinson, Bull. Am. Phys. Soc., 13, 20 (1968). ’b’h 105. G. Malli and S. Fraga, Theoret. Chim. Acta, 5, 275 (1966). m 106. D. W. Hofmeister and W. H. Flygare, J, Chem. Phys., .33' 3584 (1966). 107. F. J. Adrian, Phy . Rev., 136, A980 (1964). —_ ’b’b’b 108. C. Deverell and R. E. Richards, Molecular Physics, {$8, 551 (1966). 109. A. Saika and C. P. Slichter, J. Chem. Phys., 22, 26 (1954). mm 158 90. T. R. Tuttle, Jr., Chem. Phy . Lett., 20, 371 (1973). ————' ‘———— xx 91. R. V. Pound, Phys. Rev., 1%, 685 (1950). 92. H. S. Gutowsky and B. R. McGarvey, J. Chem. Phys., 21, 1423 (1953). 93. N. F. Ramsey, Phys. Rev., 6, 699 (1950). 94. K. Yosidu and T. Moriya, J, Phys. Soc. Japan, 11, 33 (1956). mm 95. J. Kondo and J. Yamashita, J. Phy . Chem. Solids, JR, 245 (1959). 96. D. Ikenberry and T. P. Das, Phy . Rev., 138, A822 (1965). “‘ mmm 97. D. Ikenberry and T. P. Das, J. Chem. Phys., 43, 2199 (1965). ”m 98. C. Deverell, Progr. Nucl. Magn. Reson. Spectrosc., 4, 278 (1969). 99. O. Lutz, Physics Letters, 25A, 440 (1967). — ’b'V'M 100. O. Lutz, A. Naturforschung, 23a, 1202 (1968). _ ’b’h’b 101. O. Lutz and A. Schwenk, Physics Letters, 24A, 122 (1967). “m” 102. A. Beckmann, K. D. Boklen and D. Elke, J. physik, 338, 173 (1974). 103. C. W. White, W. M. Hughes, G. S. Hayne and H. G. Robinson, Phys. Rev., 174, 23 (1968). ’Vb’h 104. G. S. Hayne, C. W. White, W. M. Hughes and H. G. Robinson, Bull. Am. Phys. Soc., 13, 20 (1968). ’\J’\: 105. G. Malli and S. Fraga, Theoret. Chim. Acta, 5, 275 (1966). m 106. D. W. Hofmeister and W. H. Flygare, J. Chem. Phys., 33, 3584 (1966). - 107. F. J. Adrian, Phy . Rev., 136, A980 (1964). ——_ mmm 108. C. Deverell and R. E. Richards, Molecular Physics, 33, 551 (1966). 109. A. Saika and C. P. Slichter, J. Chem. Phys., 22, 26 (1954). mm 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. C. Deverell, Mol. Phys., 16, 491 (1969). ’\/\1 R. H. Erlich and A. I. Popov, J. Am, Chem. Soc., 93, 5620 (1971). “m M. Herlem and A. I. Popov, J. Am, Chem. Soc., 94, 1431 (1972). “W V. Gutmann and E. Wychera, Inorg. Nucl. Chem. Lett., 2, 257 (1966). m Y. M. Cahen, J. L. Dye and A. I. Popov, J, Phys. Chem., 79, 1289 (1975). mm A. Abragam, "The Principles of Nuclear Magnetism," Oxford UP, London (1961). R. M. Sternheimer, Phys. Rev., 84, 244 (1951); 86, 316 (1952); 95, 736 (1954). mm F. W. Langhoff and R. P. Hurst, Phys. Rev., 139A, 1415 (1965). “mm“ J. P. Kintzinger, J. M. Lehn, J. Amer. Chem. Soc., 96, 3313 (1974). mm R. E. Richards and B. A. Yorke, Mol. Physics, 6, 289 (1963). m R. A. Craig and R. E. Richards, Trans. Faraday Soc., 59, 1972 (1963). ’b’b C. Deverell, D. J. Frost and R. E. Richards, Mol. Physics, 9, 565 (1965). ’b C. Hall, G. L. Haller, and R. E. Richards, Mol. Physics, 16, 377 (1969). "-—-—“' rvn C. Hall, R. E. Richards, G. N. Schulz and R. R. Shang, M01. Physics, 16, 528 (1969). mm H. G. Hertz, Ber. Bunsenges. Physik. Chem., 78, 531 (1973). ”” R. R. Dewald, Ph.D. Thesis, Michigan State University (1963). Micovic, O. S., A, 264. 127. 128. 129. 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 160 D. D. Traficante and J. A. Simms, J, Mag. Res., 33, 484 (1974). D. A. Wright, Ph. D. Thesis, Michigan State University (1974). V. A. Nicely and J. L. Dye, J. Chem. Educ., $8 (1971). P. D. Moras and R. Weiss, Acta Cryst., 829, 396, 400 (1973). “m” P. D. Moras, B. Metz and R. Weiss, Acta Cryst., B29, 383, 388 (1973). “m“. J. M. Lehn, Struc. Bonding (Berlin), Ag, 1 (1973). R. M. Izatt, D. J. Eatough and J. J. Christensen, Struct. Bonding (Berlin), 16, 161 (1973). ’b J. J. Christensen, D. J. Eatough and R. M. Izatt, Chemical Reviews, 74, 351 (1974). ’b'b Y. M. Cahen, J. L. Dye and A. I. Popov, J, Phys. Chem., 79, 1292 (1975). ’b’b Shchori, J. Jagur-Grodzinski, Z. Luz and M. Shporer, Amer. Chem. Soc., 93, 7133 (1971). mm |c4m E. Shchori, J. Jagur—Grodzinski and M. Shporer, J, Amer. Chem. Soc., 95, 3842 (1973). ———— “—— mm M. Shporer and Z. Luz, J, Amer. Chem. Soc., 97, 665 (1 75). ”“ S. Gutowsky, D. W. McCall and C. P. Slichter, Chem. Phys., 21, 279 (1953). _ ’L’b |C4m :12 M. McConnell, J, Chem. Phys., 2%, 430 (1958). R. K. Gupta, T. P. Pitner and R. Wasylishen, J, Mag. Res., 13, 383-385 (1974). *“ mm D. E. Woessner, J. Chem. Phys., 35, 41 (1961). _ ’b’b B. G. Cox, G. R. Hedwig, A. J. Parker and D. W. Watts, Aust J. Chem, 27, 477 (1974). - 'b'b W. Lamb, Phys. Rev., fig, 817 (1941). 161 145. E. Clementi, Tables 2E Atomic Functions, International Business Machines Corporation, (1965). 146. R. H. Erlich, E. Roach and A. I. Popov, J, Amer. Chem. Soc., 2%, 4989 (1970). 147. M. S. Greenberg, R. L. Bodner A. I. Popov, J, Phys. Chem, 77, 2449 (1973). ——,\J’\J 148. J. D. Haliday, R. E. Richards and R. R. Sharp, Proc. Roy, Soc. Lond., A313, 45 (1969). mmmm 149. M. R. Baker, C. H. Anderson and N. F. Ramsey, Phys. Rev., A133, 1533 (1964). ’b’b’b’b ”'TITJ'IMHMJM)(NJIQQRJHEQMWMDAS