lll'lllll/Illllllllsl llllllTl 312903 LARY Michigan State University This is to certify that the dissertation entitled Ion Association and Complexation of Alkali Cations by Crown Ethers and Cryptands in Various Solvents presented by Sadegh Khazaeli has been accepted towards fulfillment of the requirements for Ph.D. degeehi Physical Chemistry ' 2 Major professor Dam August 4, 1982 MSU is an Affirmative Anion/Equal Opportunity Instilutiou 0— 12771 MSU LIBRARIES » RETURNING MATERIALS: Place in book drop to remove this checkout from your record. FINES will be charged if boOk is returned-after the date stamped below. NM 1 51994 ION ASSOCIATION AND COMPLEXATION OF ALKALI CATIONS BY CROWN ETHERS AND CRYPTANDS IN VARIOUS SOLVENTS By Sadegh Khazaeli A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1982 ~} /'3¢'&S ABSTRACT ION ASSOCIATION AND COMPLEXATION OF ALKALI CATIONS BY CROWN ETHERS AND CRYPTANDS IN VARIOUS SOLVENTS By Sadegh Khazaeli The concentration and temperature dependence of the 13303 chemical shift, 6, for cesium salts in methylamine was fit by a model involving ion—pairs and triple—ions. Formation constants for ion—pairing, K. at 25.0°C and ip’ standard deviations are: (2.65:0.19) x 105 and (1.30i 0.19) X lOLl M_1 for CsI and CsBPhu respectively, and the corresponding enthalpies of association are: 3.7:O.3 and l. The value for (K. obtained lp)CSI from conductance measurements depended on the conductance A.O:l.0 kcal.mole_ equation used. The Onsager limiting law fit the data best A —l with (K. ) (3.u3io.2o) x 10 M at —15.7°c which is ip 051 = only about 35% of the value obtained from the NMR data at this temperature. The variation of 5 with temperature and with Sadegh Khazaeli (l8—Crown—6)/(Cs+) mole ratio (R) in methylamine and am— monia solutions indicated the formation of both 1:1 and 2:1 complexes. The concentration and temperature dependence of 6 for the 1:1 complex in methylamine was described by K K the equilibria: Cs+ + C‘-f Csc+, Csc+ + X‘ :A Csc+.x‘, ARC . AHX K +_ X +— _ and Cs .X -+C EH0 CsC .X , where C and X are the ligand x and the anion respectively. Values of KC = (1.07:0.08) x 10L1 M‘1 at 25.000 and AH: = —l6.72i0.08 kcai.moie‘l were obtained. The ion—pair formation constants of the 1:1 complex at 25.0°C are: (KA)CSI = (l.51i0.06) x 105 and (KA)CsBPhh = (1.16:0.34) x 10A M—l. Other parameters have the values: KX = (8.4ii.u) x 103, (6.33:0.uo) x 103, and (4.87:0.53) x 103 M‘1 at 25.000 with AH; = —l8.80i0.95, —16.uoio.53, and —13.50i0.73 keg-iihoie‘l for CsBPhu, CsI, and CsSCN, respectively.' The mole ratio'data for R > 1 were analyzed for solu- . . . + _ tions in methylamine according to the equilibrium CsC .X K + C A§§2 CEC:.X_ with corrections introduced for ion—associa- x2 tion of the salt and the 1:1 complex. The thermodynamic parameters are: (Kx2)CsI = “-03:0'05’ and (Kx2)CsBPhh = 22.82i0.35 M“1 at 25.000 with (AH;2)C8I = —6.05i0.08, and < = —7.35:O.l2 kcal.mole_l. The corresponding 0 AHx2)CsBPhh approximate values in ammonia are: (Kx2)CsBPh4 = 649i44 M“1 at 25.000 with (AHO = -u.9i:o.28 keai.moie‘l. x2)CsBPhu Rubidium-87 chemical shifts of some rubidium salts were Studied as a function of concentration. Attempts to use ¥ Sadegh Khazaeli 87Rb NMB to study complexation of the Rb+ ion by l8—crown—6 and cryptand-222 failed. The exchange of the Li+ ion between the free and 211— cryptated species in methylamine solutions, while slow on the NMR time scale, occurs within seconds of mixing. The chemical shift of Li+ in the complex is the same in methyl- amine and ammonia, indicating that the complexant effec— tively isolates Li+ from the solvent and anion. To my wife and in memory of my father ii - ' ' 3 ~77 - ' ' _. v‘L. "~ “(1’35 ;’j.."‘". T ..- _ ‘T-‘mf. 7.. A . ACKNOWLEDGMENT The author wishes to express his deepest appreciation to Professor J. L. Dye for his encouragement, enlighten— ing suggestions and invaluable assistance, and to Professor A. I. Popov for his guidance, whole-hearted support, and friendship throughout the course of this investigation. Gratitude is extended to the Department of Chemistry, Michigan State University, and the National Science Founda— tion for financial support under Grants DMR-79—2lA79 and CHE—80-lO808. Thanks are also extended to the members of the research groups of both Professor Popov and Professor Dye for their help and moral support, and to the friends at Michigan State University and elsewhere without whose friendship this study would have been more difficult. Many thanks go to MSU's technical and clerical staff especially Mr. Wayne Burkhardt and Mr. Tomas Clarke for the maintenance of the NMR spectrometers, Mr. Keki Mistry for his excellent glassblowing service and secretary Sharon Corner for her assistance. Above all, the author wishes to thank his family and his wife's family for their unending encouragement, under— standing, and financial aid. I am deeply grateful to my wife, Afsaneh, for her love, patience, and encouragement through the years of graduate study and for her constant care of our sons, Javad and Nima. To her and in memory of my father, I dedicate this thesis. iv (Iv. ...-~ - '0.- --» . TABLE OF CONTENTS Chapter Page LIST OF TABLES. . . . . . . . . . . . . . . . . . . xi LIST OF FIGURES . . . . . . . . . . . . . . . . . . xx CHAPTER I - INTRODUCTION AND HISTORICAL . . . . . . I Introduction. . . . . . . . . . . . . . . . . . 2 Historical Background 3 1. Ion Association 3 A. Theoretical Aspects 3 B. Methods of Studying Ion- Association . . . . . . . . . 1A i. ESR, Electronic, and Vibra- - tional Spectroscopy . . . . . . 15 a. ESR spectroscopy. . . . .i. 15 b. Electronic spectroscopy . . 16 c. Vibrational spectroscopy. . 18 ii. Electrical Conductance Measure- ment. . . . . . . . . . . . . . 20 iii. Alkali Metal Nuclear Magnetic Resonance Spectroscopy. . . . . 27 a. Introduction. . . . . . . . 27 b. Relaxation Studies. . . . . 28 0. Chemical Shift Studies. . . 31 2. Complexation of Alkali Cations by Crown Ethers and Cryptands . . . . . A5 A. Introduction. . . . . . . . . . . . A5 B. Selectivity of Complexation . . . . A9 1. Relative Sizes of Cation and Ligand Cavity . . . . . . . . . U9 .'. .’. Chapter ii. Arrangement of Ligand Binding Sites . iii. Type and Charge of Cation iv. Type of Donor Atom. v. Number of Donor Atoms vi. Substitution of the Macro— cyclic Ring Vii. Solvent and Anion Effects C. Thermodynamics of Complexation. D. The Use of Alkali Metal NMR to Study Thermodynamics and Kinetics of the Complexation. CHAPTER II — EXPERIMENTAL 1. Purification of Materials A. Ligands B. Solvents. C. Salts 2. Glassware Cleaning. 3. NMR Techniques. A. NMR Instruments B. Data Handling C. NMR Sample Preparation. . . . A. Conductance Method. A. Conductance Equipments. B. Data Handling C. Sample Preparation for Conductance Measurements. . . . . . . . . CHAPTER III — THERMODYNAMICS OF ION—ASSOCIATION OF CESIUM SALTS IN METHYLAMINE. vi Page 53 53 55 55 56 57 58 61 6A 65 65 67 69 70 7O 7O 72 75 8O 8O 81 83 85 ‘ nu. '.‘.~.‘ . . A _: a“!' ”1. n a . o o ‘1' Chapter Introduction. Cesium-133 NMR Studies of Cesium Salts in Methylamine. . . . A. Concentration Dependence of 133Cs Chemical Shifts of Cesium Salts in Methylamine. . . . . . . B. Ion Association of Cesium Salts in Methylamine. i. Simple Ion—Pair Model ii. Formation of Triple—Ions. a. One type of triple-ion. b. Two types of triple-ions. C. Complimentary Experiments i. Cesium Tetraphenylborate in 90% V/v Methylamine in Dimethyl- sulfoxide Solutions . . ii. Mixtures of Cesium Iodide and Cesium Thiocyanate in Methyl— amine . . iii. Mixtures of Cesium Iodide and Cesium Tetraphenylborate in Methylamine D. Conclusions Electrical Conductance Measurements of Cesium Iodide in Methylamine A. Introduction. B. Results 1. Calibration of the Conductance Cell. . . ii. Conductance of Cesium Iodide in Methylamine at -15.7°C. C. Discussion. D. Conclusion. vii Page 86 87 87 103 103 108 108 110 129 130 133 137 1A3 1A5 1A5 1A6 1A6 149 152 155 Chapter Page A. Comparison of NMR and Electrical Conductance Measurements. . . . . . . . . . 156 CHAPTER IV — COMPLEXATION OF CESIUM SALTS BY 18-CROWN-6 IN METHYLAMINE AND LIQUID AMMONIA . . . . . . . . . . . . 159 1. Introduction. . . . . . . . . . . . . . . . 160 2. Complexation of Cesium Salts by 18—crown-6 in Methylamine . . . . . . . . . 161 A. Mole Ratio Dependence of 13303 Chemical Shift in Methylamine . . . . . 161 i. Results . . . . . . . . . . . . . . 161 ii. Discussion. . . . . . . . . . . . . 169 B. Concentration Dependence of the 133Cs Chemical Shift of the 1: 1 Complex in Methylamine. . . . . . . 176 i. Results . . . . . . . . . . . . . . 176 ii. Discussion. . . . . . . . . . . . . 183 a. General discussion. . . . . . . 183 b. The chemical shift of the 1:1 complex, 5MC+ . . . . . . . 185 c. Enthalpy of formation of the ion—paired complex. . . . . . . 189 iii. Analysis of the Data. . ; . . . . . -191 a. Cesium iodide . . . . . . . . . 191 b Cesium tetraphenylborate. . . . 194 c. Cesium thiocyanate. . . . . . . 197 d Summary . . . . . . . . . . . . 201 C. Thermodynamics of Formation of 2: 1 Complexes of 18- crown-6 With Cesium Salts in Methylamine. . . . . . . . . . 205 3. Complexation of Cesium Tetraphenyl— borate by 18— crown-6 in Liquid Ammonia. . . 221 A. Results . . . . . . . . . . . . . . . . 221 viii '.-.--;fl I . up. ‘0 q I ‘0 '\ i“ .-' -‘ e ‘ .~. '§ ‘\ e .. i. .' ‘5 . ‘p. C‘ I . ._. I h 'n | V O II, It Chapter Page B. Discussion. . . . . . . . . . . . . . . 226 C. Summary . . . . . . . . . . . . . . . . 231 A. Conclusion. . . . . . . . . . . . . . . . . 233 CHAPTER V — l. RUBIDIUM-87 NMR INVESTIGA- TION OF RUBIDIUM SALTS AND THEIR COMPLEXES IN AQUEOUS AND NON- AQUEOUS SOLVENTS 2. LITHIUM—7 NMR STUDY OF COM- PLEXATION OF LITHIUM SALTS BY C211 IN METHYLAMINE AND LIQUID AMMONIA . . . 237 1. Introduction. . . . . . . . . . . . . . . . 238 2. Investigation of Rubidium Salts and Their Complexes in Aqueous and Non- Aqueous Solvents ....... . . . . . . . . . . 239 A. Salt Solutions. . . . . . . . . . . . . 239 B. Complexation. . . . . . . . . . . . . . 2AA 3. Complexation of the Lithium Cation by C211 in Methylamine . . . . . . . . . . . . 2A8 A. Complexation of the Lithium Cation by C211 in Liquid Ammonia. . . . . . . . . . . 252 5. Conclusion. . . . . . . . . . . . . . . . . 252 CHAPTER VI - SUMMARY AND SUGGESTIONS FOR FURTHER STUDIES. . . . . . . . . . . 255 1. Summary . . . . . . . . . . . . . . . . . . 256 2. Suggestions for Further Studies . . . . . . 261 APPENDICES. . . . . . . . . . . . . . . . . . . . . 26A APPENDIX 1 - DETERMINATION OF ION-ASSOCIATION PARAMETERS BY NMR TECHNIQUES: DES— CRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN. . . . . . . 265 A. Simple Ion—Pair Formation . . . . . . . . . 265 B. Ion-Pairs and Anionic Triple-Ion Formation . . . . . . 267 0- Ion-pairs and Two Kinds of Triple Ions. . . 27A ix Chapter APPENDIX 2 - DETERMINATION OF ION-PAIR FORMA— TION CONSTANTS BY CONDUCTANCE MEASUREMENTS, DESCRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUB- ROUTINE EQN. . . . A. The Onsager Limiting Law. B. Extended Conductance Equation i. Pitt's Equation Linearized by Fernandez-Prini ii. Fuoss-Hsia Equation Linearized by Fernandez Prini. iii. Fuoss-Hsia Equation Corrected by Chen iv. Justice Equation. APPENDIX 3 - DETERMINATION.OF COMPLEX FORMATION CONSTANTS BY THE NMR TECHNIQUE; DESCRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN. A. 1:1 Complex Formation in Media of Low Dielectric Constant B. 2:1 Complex Formation in Media of Low Dielectric Constant REFERENCES. Page 28A 28A 286 290 291 298 298 302 302 315 32A _._—.._ . . _........._._..__..... .’~-. Table LIST OF TABLES Page Nuclear Properties of Alkali Nuclei . . . . . 29 Chemical Shifts/ppm at Infinite Dilution for 7Li+, 23Na+, 39K+ 1330s;+ Ions in Various 801- and vents. . . . . . . . . . . . . . . . . . . A2 Diamagnetic Susceptibility Correction of Various Solvents on DA-60 and WH-180 Instruments . . . . . . . . . . . . . . . . 7A 133Cs Concentration Dependence of Chemical Shift of CST in Methylamine at Various Temperatures . . . . . . . . . . 88 Concentration Dependence of the 13305 Chemical Shift of CsBPhu in Methyl- amine at Various Temperatures . . . . . . . 91 Variation of the 13305 Chemical Shift with the Mole Ratio (18C6)/ (CsBPhu) in Methylamine Solutions at Different (Cs+) Concentrations and Various Temperatures. . . . . . . . . . . . 93 Concentration Dependence of the 13305 Chemical Shift of CsSCN in Methylamine at Various Temperatures . . . . . . . . . . 95 xi ‘Table 10 11 12 13 Concentration Dependence of the 133CS Chemical Shift of Csi3 in Methylamine at 25.00 and 5.6°C. Cesium-133 Chemical Shifts of Saturated Solutions of CsBr, CsClOu, CsNO3 in Methylamine at 25.000. Calculated Thermodynamic Parameters of Ion-Association of Cesium Salts in Methylamine at 25.0°C According to a Simple Ion—Pair Model Calculated Thermodynamic Parameters for Ion-Association of Cesium Salts in Methylamine at 25.000 According to the Ion-Pair and Anionic Triple—Ion Model Thermodynamic Parameters for Ion— Association of Cesium Salts in Methyl— amine at 25.0°C; Obtained with the Assumptions that Ion-Pairs and Two Kinds of Triple Ions are Present, and That 50s = 60.73 ppm at 25.000. Thermodynamic Parameters for Ion- Association of Cesium Salts in Methyl- amine at 25.000; Obtained with the Assumptions that Ion-Pairs and Two xii Page 98 100 107 111 117 Table 13 1A 15 16 17 Kinds of Triple Ions are Present, and That 6C3 = A0 ppm Thermodynamic Parameters for Ion-~ Association of Cesium Salts in Methyl- amine at 25.000; Obtained With the Assumption that Ion-Pairs and Two Kinds of Triple Ions are Present, and That 603 = 80 ppm Thermodynamic Parameters for Ion— Association of Cesium Salts in Methyl- amine at 25.0°C; Obtained with the Assumptions that Ion-Pairs and Two Kinds of Triple Ions are Present, and That Kt/Kip is a Known Constant Thermodynamic Parameters for Ion- Association of Cesium Salts in Methyl- amine at 25.0°C; Obtained with the Assumptions that Ion-Pairs and Two Kinds of Triple-Ions are Present, and that Kt/Kip is an Adjustable Parameter Thermodynamic Parameters for Ion- Association of Cesium Salts in Methyl- amine at 25.0°C; Obtained with the Assumptions that Ion-Pairs and Two xiii Page 118 119 122 12A Table 17 18 19 20 21 22 Page Kinds of Triple Ions are Present, with Equipartition of the Chemical Shifts. . . . . . . . . . . . . . . . . . . 127 Thermodynamic Parameters for Ion- Association of Cesium Salts in Methyl- amine at 25.0°C; Obtained with the Assumptions that Ion-Pairs and Two Kinds of Triple Ions are Present, and is Temperature Dependent . . . . 128 133Cs that 60$.X Concentration Dependence of the Chemical Shift of CsBPhu in 90% v/v Methylamine-DMSO at 25.0°C. . . . . . . . . 131 133Cs Chemical Shifts of Different Mole Ratios of (I-)/(SCN-) + (I-) in Mixtures of CsI and CsSCN in Methylamine at Various Temperatures; (Cs+) = 0.005A E- . . . . . . . . . . . . . 13A Cesium-133 Chemical Shifts of Dif- ferent Mole Ratios of (I-)/(BPhE) + (I‘) for 0.0007 and 0.003 g 0s+ in Methylamine Solutions at 2A.6°C . . . . . . 138 Resistances of the 90.9k0 Standard Resistor, Water, and Aqueous Potas- sium Chloride Solutions at 25.0:0.0A°C. . . 1A7 xiv Table 23 2A 25 26 Equivalent Conductances of Cesium Iodide Solutions in Methylamine at —15.7°C . . . . . Values of the Association Constant and Limiting Equivalent Conductance of Cesium Iodide in Methylamine at —15.7°C Obtained by Various Conductance Equations Mole Ratio Study of 18C6, CsI Complexes in Methylamine at Various Temperatures; (08+) = 0.0206i0.0008 1_v1_ . Mole Ratio Study of 1806, CsBPhu Complexes in Methylamine at Various Temperatures; (Cs+) = 0.0108i0.0005 M Thermodynamic Parameters for the Formation of the 2:1 Complex of 18- .Crown—6 and CST in Methylamine. K i 1 10M Thermodynamic Parameters for the Forma— tion of the 2:1 Complex of 18—Crown-6 and CsBPhu in Methylamine. Kl : 10M Concentration Dependence of the 133Cs Chemical Shift of the 1:1 Complex of CsI and 18C6 in Methylamine at Various Temperatures. XV Page 150 15A 162 Table 30 31 32 33 3A 35 Page Concentration Dependence of the 13305 Chemical Shift of the 1:1 Complex of CsBPhM and 18C6 in Methylamine at Various Temperatures. . . 178 Concentration Dependence of the 133Cs Chemical Shift of the 1:1 Complex of CsSCN and 1806 in Methylamine at ‘ Various Temperatures. . . . . . . . . . . . 179 Thermodynamic Parameters for the. Formation of the 1:1 Complex of 18- Crown—6 and CsI in Methylamine at 25.000; with Adjustment of AHZ. . . . . . . 190 Thermodynamic Parameters for the Formation of the 1:1 Complex of 18C6 and CsI in Methylamine at 25.000; with Calculation of AH: from Other Adjustable Parameters. . . . . . . . . . . . . . . . . 193 Thermodynamic Parameters for the Formation of the 1:1 Complex of 18C6 and CsBPhu in Methylamine at 25.000; with AH: = Ang . . . . . . . . . . . . . . 195 Thermodynamic Parameters for the Formation of the 1:1 Complex of 1806 and CsBPhu in Methylamine at 25.000; KC was Used as a Constant . . . . . . . . . 198 Thermodynamic Parameters for the Formation of the 1:1 Complex of 18C6 and CsSCN in Methylamine at 25.0°C. Thermodynamic Parameters of the Complexation of Cesium Salts by 18— Crown—6 in Methylamine at 25.0°C. Thermodynamic Parameters for the Formation of the 2:1 Complex of CsI with 18—Crown-6 in Methylamine at 25.000; Assuming Complete Ion-Pair Formation of the MC+ Complex at Thermodynamic Parameters for the Formation of the 2:1 Complex of CsI with 18—Crown-6 in Methylamine at 25.0°C; Assuming Complete Ion—Pair Formation of Both the MC+ and MC+ Complexes at (Cs+) Thermodynamic Parameters for the Formation of the 2:1 Complex of CsBPhu with 18—Crown—6 in Methylamine at 25.0°C; Assuming Complete Ion—Pair Formation of both the MC+ and MC+ Complexes at (Cs+) Page 200 202 207 210 211 Table A1 A2 A3 AA A6 A7 -Concentration Dependence of the Thermodynamic Parameters for the Formation of the 2:1 Complexes of Cesium Salts with 18-crown-6 in Methylamine at 25.0°C . . . 133CS Chemical Shift of CST in the Presence of a 6.0-Fold Excess of 18—Crown—6 in Methylamine at Various Temperatures. Mole Ratio Study of 18C6.CsBPhu Com— plexes in Liquid Ammonia at Various Temperatures; (Cs+) = 0.001 M . Mole Ratio Study of 18C6.CsBPhu Com— plexes in Liquid Ammonia, at Various Temperatures; (08+) = 0.0075 M. 133CS Concentration Dependence of the Chemical Shift of CsBPhuin Liquid Ammonia at 6.0°C. Complexation Formation Constant, Limiting Chemical Shifts, and Thermo- dynamic Parameters for the 2:1 Complex of 1806, CsBPhu in Liquid Ammonia at Various Temperatures, Assuming K1 1 10A Thermodynamic Parameters for the Formation of the 2:1 Complex of 1806 and Cesium Salts in Methylamine and, Ammonia Solutions . . . . . . . . . . xviii Page 216 219 222 225 227 232 Table - Page A8 Rubidium—87 Chemical Shifts and Linewidths of Rubidium Salts in Three Solvents at Ambient Temperatures. . . . . . 2A0 A9 Rubidium~87 Chemical Shifts versus (Ligand)/(Rb+) Mole Ratio . . . . . . . . . 2A5 50 Lithium—7 Chemical Shifts of the Free, 5F, and C211—Cryptated, 0C, Lithium Cation in Methylamine as a Function of Time. (211)/(Li+) = 0.5, (Li+) = 0.069 M . . . . . . . . . . . . . . . . . 2A9 7Li Chemical 51 Mole Ratio Study of the Shift of LiBr in the Presence of Cryptand 211 in Liquid Ammonia. (Li+) = 0.07 g. . . . . . . . . . . . . . . 253 LIST OF FIGURES Figure Page 1 Cesium—133 chemical shift as a function of concentration. . . . . . . . . 35 2 Cesium-133 shifts against concentra— tion for CsBr+HOAc-H20 systems . . . . . :7 "38 3 Sodium-23 chemical shifts YE Gutmann donor numbers. . . . . . . . . . . . . . . A0 A Naturally occurring and synthetic macrocycles. . . . . . . . . . . . ... . . A6 5 (a) Various stoichiometries of K+-crown ether complexes. (b) Exclusive CS 0222 complex . . . . . . . . . . . . . . . ' A8 6 Selectivity of 18-crown—6: log K values for the reaction of 18—crown—6 with metal cations in H2O YE: ratio of cation diameter to l8—crown—6 cavity diameter .‘. . . . . . . . . . . . . . . . 51 7 Selectivity of cyclic polyethers of various sizes: log K values for re— action of several crown ethers with alkali metal ions K§.cation radius . . . . 52 XX Figure 10 Ill :12 3.3 3.4 :1 1-7' 55 Selectivity of cryptands: log K values for reaction of several cryptands with alkali metal cations gs. cation radius Cryptand sublimation apparatus Calibration curve for 10 mm NMR tubes at 0°C . Extended NMR tube for high vacuum. Extended NMR tube for the kinetic ex- periment The conductance cell . . Concentration and temperature depend- ence of the 133Cs chemical shift of cesium iodide in methylamine Concentration and temperature depend- ence of the 133CS chemical shift of cesium tetraphenylborate in methyl— amine. Cesium-133 chemical shift versus (18-crown-6)/(CSBPhu) mole ratio at different concentrations of CsBPhu and various temperatures . Concentration dependence of the 13303 chemical shift of cesium thiocyanate in methylamine at various tempera; tures. xxi Page 5A 66 76 77 79 82 90 92 9A 96 Figure 18 19 220 2323 214 255 Concentration dependence of the 133Cs chemical shift of cesium triiodide in methylamine at 25.0 and 5.6°C. Concentration dependence of the 133CS chemical shifts of some cesium salts in methylamine at 25.0°C Concentration dependence of 13303 chemical shifts of CsI and CsBPhu in methylamine at 25.0°C Cesium-133 chemical shifts of cesium thiocyanate solutions versus tempera- ture Concentration dependence of the 133Cs chemical shift of cesium tetraphenyl- borate in 90% v/v methylamine in di- methylsulfoxide at 25.0°C. A comparison of 133Cs chemical shifts of pure cesium salts and the mixture of cesium salts in methylamine at 25.0°C Plot Of (5obs‘50sl)/(50sSCN‘50sl) versus mole fraction of iodide A comparison of 133CS chemical shifts of pure cesium salts and mixtures of cesium salts in methylamine at 25.0°C. xxii Page 99 101 106 116 132 135 136 139 Figure 26 27 £28 2259 3ZL 322 333 Plot of (Gobs'GCsI)/(SCsBPhu-SCSI) versus mole fraction of iodide Calculated 13305 chemical shifts of mix- tures of CsI and CsTPB as a function of concentration Plot of equivalent conductance versus the square root of the molar concentra- tion of cesium iodide in methylamine at ~15.7°C Cesium-133 chemical shift versus (l8-crown-6)/(CsI) mole ratio and temperature in methylamine; (CsI) = 0.02 M . Cesium—133 chemical shift versus (l8-crown-6)/(CsBPhu) mole ratio and temperature in methylamine; (CsBPhu) = 0.01 M. Ln"K2" XE l/T for the 2:1 complex of 18-crown—6 and CsI in methylamine. Kl Z 10”, (C31) = 0.02 M . Ln "K2" gs l/T for the 2:1 Complex of 18-crown-6 and CsBPhu in methylamine. A K1 3 10 , (CsBPhu) = 0.01 g. Concentration dependence of the 133Cs chemical shift of the 1:1 complex of XXiii Page 1A1 1A2 151 16A 168 173 17A Figure 33 3A 35 36 37 38 39 7' "' W. ,' _ '..' ' —: WEI—m —' 18-crown-6 and CsI in methylamine at various temperatures . . . . Concentration dependence of the 1330s chemical shift of the 1:1 complex of 18-crown—6 and CsBPhu in methylamine at various temperatures . . . . Concentration dependence of the 133Cs chemical shift of the 1:1 complex of 18—crown—6 and CsSCN in methylamine at various temperatures Limiting 1330s chemical shift of the 1:1 complex of 1806 and CsSCN XE tem— perature in methylamine. . . . . . . . Concentration dependence of the 133Cs chemical shift of CST in the presence of a 6.0—fold excess of lB—crown—6 in methylamine at various tempera- tures. . . . . Cesium—133 chemical shift versus (18—crown—6)/(CsBPhu) mole ratio and temperature in liquid ammonia; (Cs+) = 0.001 g. . . CeSium—133 chemical shift versus (18—crown—6)/(CsBPhu) mole ratio in liquid ammonia at various tem— Page 180 181 182 187 220 . . . 223 peratures; (Cs+) = 0.0075 M, . . . . . . . 22h xxiv Figure A0 A1 U42 Page Cesium-133 chemical shift versus concentration of CsBPhu in liquid ammonia at 6.0°C . . . . . . . .l. . . . . 228 Rubidium-87 chemical shifts of rubidium bromide versus concentration (—) or mean molar activity (---) in aqueous solutions. . . . . . . . . . . . . . . . . 2A2 Rubidium-87 chemical shifts of rubidium bromide in methanol (-) or rubidium iodide in propylenecarbonate (-—-) versus concentration (0) or mean molar activity (A) of the solution . . . . . . . 2A3 XXV CHAPTER I INTRODUCTION AND HISTORICAL 1 INTRODUCTION Since the discovery of crown ethers by Pederson (1—3) and cryptands by Lehn and coworkers (A-7), the thermo— dynamics and kinetics of complexation of alkali cations by these novel compounds have been studied in aqueous and nonaqueous solvents by various physicochemical techniques. Most of the research in this area has been performed in relatively high dielectric media, where ionic association occurs to a limited degree if at all. Little is known about the effect of different ion—association processes upon complexation in low dielectric solvents. The high sensitivity of 133Cs NMR along with its narrow lines al— lowed us to investigate extensively the thermodynamics of the complexation of cesium salts by l8—crown—6 in the low dielectric constant solvent methylamine. The results of this study are presented in Chapters III and IV of this Dissertation and are compared with the data in liquid am— monia and in other solvents. The thermodynamics of complexation of all the alkali cations except the rubidium cation by crown ethers and cryptands has been studied by the alkali metal NMR tech— nique. To fill the gap, attempts were made to use 87Rb NMR to study the complexation of rubidium salts in solu— tion. This work will be described in Chapter V. The kinetics of complexation of alkali cations with crown ethers and cryptands has been investigated mostly by line shape analysis of NMR signals. It was suggested that the rate of exchange of Li+ ion between the solvated and C211 cryptated states in methylamine is very slow the exchange rate can be determined directly Chapter (8). If so, from the time dependence of the 7Li chemical shift. V also describes the kinetic study of the complexation of the lithium cation by C211 in methylamine. HISTORICAL BACKGROUND 1. Ion Association A. Theoretical Aspects Michael Faraday (9) was the first to realize that ions play an important role in the behavior of electrolytic solutions. He assumed, however, that in solutions ions are formed only under the influence of an electrical field.» Arrhenius, after extensive studies on the conductivity of electrolytic solutions, proposed that molecules of an electrolyte are spontaneously dissociated, to a certain extent, into free ions (10). His theory could explain the behavior of weak electrolytes where only a small frac— tion of the molecules are dissociated but the behavior of strong electrolytes shows considerable deviations from Arrhenius' theory. The concept of "ionic atmosphere" which is characterized by a certain regularity in the distribution of ions and the use of Poisson's equation by Debye resulted in the Debye- Hfickel limiting law (11) (DHLL) for the activity coefficient of ions.. The DHLL describes the behavior of strong elec- trolytes only at very low concentrations which is a conse- quence of the approximation used to linearize the Poisson— Boltzmann distribution function. A modification of the DHLL considers ions with finite size instead of point charges by introducing an ionic parameter "a" which is the minimum average distance to which two ions can ap— proach each other. The Debye-Hfickel theory was successful in describing experimental data at low concentration (<0.01 M) for 1:1 electrolytes in high dielectric constant solvents. At biggier concentrations, and for ions with high charges, the lirnsarized Poisson-Boltzmann equation is not valid, and true short—range forces between the ions also becomes im- portant. Grunwall, LaMer, and Sandved (12) improved the Debye- Pfiickel theory by using higher order terms in the Poisson- Ekoltzmann equation. Their treatment predicts departures iflwmlthe Debye-Hackel theory for electrolytes of high ‘falence type, for electrolytes whose "a" parameters are 53nm11,2u1d for electrolytes in solvents with low dielectric clonstants. However, their treatment has been criticized because of internal mathematical inconsistencies. Although the theory of Debye-HUckel including the ion size parameter is very successful in describing many experi— mental data (13,1A), there remain some unsolved problems such as the question about the validity of the linearized Poisson-Boltzmann equation and the consideration of short— range forces between the ions. It also does not seem to be possible to extend this theory to more general models nor to higher concentrations (m0.5 M). For these reasons much theoretical research has been done on the basis of the exact statistical treatment of the problem (15—18). A complete solution of the Poisson—Boltzmann equation is not possible due to mathematical difficulties. Bjerrum, who realized such difficulties, suggested a much simpler approach by introducing the concept of ion—pair forma— tion (19). He proposed a model of rigid nonpolarizable spheres in a medium of fixed macroscopic dielectric constant (primitive model). The ion—solvent interaction as well as nonpolar bonds were neglected. The probability that an ion is at a distance r from another ion was given by the Maxwell—Boltzmann distribution function. The Bierrum distance q is defined as the distance at which the energy of interaction of the ions is 2kT. Two ions within a distance q 1 r 2 a are considered as ion—pairs. Within this region the potential energy of the two ions was taken as the Coulombic energy. The probability of finding two 7l""""'[_ 6 - ' 7' l ions 1 and 2 in this region is, o [Z1Z2le2/DrkT 2 e r dr probability = 1 — a 1000 (l—l) where d is the degree of dissociation of the ion—pair, N is Avogadro's number, C is the molar concentration, 4 Z1 and 22 are the valences of the ions, D is the bulk di— electric constant, k is the Boltzmann constant, and T is the temperature in OK. If y E IZlZ2Ie2/erT and b IZlZ21e2/aDkT, then it can be shown that, 2 IZlZ2le 3 fb y -A _ AHNC l ’ a ' 1000 DkT 2 e y dy 2 — ””NC ’lezle >3 0(b) <1—2) — 1000 DkT b where Q(b) E f e y dY- 2 The ion-pair formation constant KA can be written as, l-d (1—3) l I -—:—-———-——-—“mr gLA'fi_‘- _.‘... . .. . I 7 | where r+ is the mean activity coefficient. As concentration decreases the mean activity coefficient and a2 approach unity, then, Introducing a from Equation (1—2) gives. [le2le2)3 KA =1‘0‘0‘0 ( DKT Q(b) (1—5) Values of Q(b) have been tabulated by Bjerrum for the range 1 3 b i 15. Values of Q(b) in the range 15 g b 5 80 were obtained by Fuoss and Kraus (20). Despite the mathematical artifacts and the oversimplification of the model, the Bjerrum theory has been able to describe systems of 1:1 electrolytes (20) as well as polyvalent electrolyte systems (21) in water and mixed solvents. One of the problems of the Bjerrum theory was that the probability function diverges for r > q and so cannot have any physical meaning. Furthermore, at low dielectric constants the value of q becomes extremely large. Fuoss (22) in 1958 defined ion-pairs to include only those ions in physical contact. The solvent was taken as a continuum dielectric and ions as rigid spheres. Upon formation of an ion-pair, the solvent molecules between the ions were squeezed out. Cations were taken as conducting spheres of radius "a" and the anions as point charges. This is equivalent to considering both cations and anions as spheres of radius a/2. By a statistical treatment the following expression for the ion—pair formation constant was ob— tained, 1—d K = A d2Cy - _ AiTNa3 b _ KA — 3000 e (l 6) The value of "a" corresponding to the minimum in KA can be obtained by differentiating the above equation with re— spect to "a". This yields, 2 — 'le2le - g <1-7> amin-W-sq The Fouss expression consists of two terms (23). The pre— eXponential factor has an entropic nature and stands for a probability factor which increases with the Size of the C01liding ion. When 'legl = 0 the value of KA is a measure Of the excluded volume from the solution occupied by the ions. The exponential term is energetic in nature and I . A _3_ expresses a shielding factor caused by the ionic charges. The Fouss theory considers the solvent to be a continuum dielectric and includes only electrostatic forces between the ions. Fouss and Kraus (2A,25) obtained an expression for the formation constant of triple—ions, by assuming equal prob- ability of the formation of cationic and anionic triple— ions. Their final equation for the case in which the centers of the three ions lie on a straight line is, 3 ~ 311: I N+Butu > Cs+ > K+ > Na+ > Li+. A similar trend was observed by Zaugg and Schaefer (55) in the spectra of alkali enolates and alkali phenolates. The existence of different kinds of ion-pairs can be detected by optical spectroscopy. Changes in solvent polarity and struCture can affect the equilibrium between contact and solvent separated ion—pairs. The absorption bands caused by solvent separated ion—pairs are often in— dependent of the cation and solvent, but the absorption bands of contact ion—pairs usually show a shift to higher wavelength as the solvent polarity increases (5A). The equilibrium between the two kinds of ion-pairs also depends on temperature, pressure, and anion structure. For ex- ample, 9—(2—hexy1)fluorenyllithium in 2,5—dimethyltetra— hydrofuran solution is predominantly in the form of contact ion pairs at -20°C, but in the form of solvent separated ion—pairs at —A0°C (5A). Szwarc and coworkers showed that for fluorenyl equilibrium 3 pressure incr on the anion solvent separ Fox and Hz the position < tion bands of for solvated i tahnd a linea electron and v 311110113 (58) ha tion. (0) V found a band i salts in benze Edgell and cow bands of alkal sulfoxide. The A the nature 0 anion. It was 3180193. POpOV of “Age numb A? [R bands We 18 for fluorenyllithium and sodium salts in tetrahydrofuran, the equilibrium shiftstowanS solvent separated ion-pairs as the pressure increases (56). The presence of large substuents on the anion also shifts the equilibriumtowards formation of solvent separated ion—pairs. Fox and Hayon (57) found a linear correlation between the position of the absorption maxima of the first absorp- tion bands of bromide and chloride ions and Vmax (cm—l) for solvated iodide in the same solvent. They also ob— tained a linear relationship between Vmax of the solvated electron and Vmax of the solvated iodide in 38 solvents. Symons (58) has reviewed uv spectroscopic studies of solva— tion. (c) Vibrational Spectroscopy — Evan and Lo (59) found a band in the far IR spectra of tetraalkylammonium salts in benzene which was assigned to ion—aggregates. Edgell and coworkers (60,61) observed similar far IR broad bands of alkali solutions in tetrahydrofuran and dimethyl— sulfoxide. The frequencies of these bands depend largely on the nature of the cation and, to some extent, on the anion. It was suggested that the bands are due to the cation—anion Vibration of ion-pairs or higher aggregate Species. Popov and coworkers (62) obtained far IR spectra of a large number of alkali salts in dimethylsulfoxide. Far IR bands were observed which could not be assigned to either solve were strongly of the anions tion of the c of 6Li+, NDZ, indicated that vibration. Popov and spectrum of 11' mixtures. The chlorate ion 8 acetone/Li+ mo broadens and SI behavior Shows l. The effect on the far IR were investiga human and IR d the existence 0 m the anion i that NaCo ( co ) A contact ion-pai th e I‘l’chofco)!l S lhethoxyethane 19 either solvent or salts. The frequencies of the bands were strongly cation dependent, but completely independent of the anions (63). These bands were assigned to the vibra- tion of the cation in a solvent cage. Isotopic substitutions of 6Li+, NDZ, d6DMSO for 7Li+, NHZ, and DMSO, respectively indicataithat both cation and solvent participate in the vibration. Popov and coworkers (64,65) studied the mid infrared spectrum of lithium perchlorate in nitromethane—acetone mixtures. The symmetric stretching vibration of the per— chlorate ion shows a narrow and symmetrical band when the .+ . acetone/Ll mole ratio is greater than M. This band broadens and shifts with decreasing mole ratio. This behavior shows that the solvation number of Li+ is M. The effects of cation, anion, solvent, and concentration on the far IR band of NaCo(CO)u in tetrahydrofuran solution were investigated by Edgell at al. (60,61). The additional Raman and IR data for the Co(CO); anion was explained by the existence of different sites (or kinds of ion—pairs) for the anion in tetrahydrofuran. The authors concluded that NaCo(CO)u exists mainly as solvent separated and contact ion-pairs (66—68). Addition of cryptand C221 to the NaCo(CO)u solutions (69) in tetrahydrofuran and 1,2— dimethoxyethane solutions converts the mix of ion sites to a single solvent separated ion site of tetrahedral symmetry. In tetrahydI’O site of 02v 3 is some inter sodium cation Popov (70 of electrolyte (ii) Elec Arrhenius behavior of we sociation of a measured equiv; and A0 is the e At low concent equilibrium AB equation (72) , The Oswald (111 20 In tetrahydropyran and 2-methyl tetrahydrofuran a new site of 02v symmetry formed. This indicates that there is some interaction between the anion and the cryptated sodium cation. Popov (70) and Irish (71) reviewed IR and Raman studies of electrolytes. (ii) Electrical Conductance Measurement Arrhenius (10) was the first to explain the conductance behavior of weak electrolytes by defining the degree of dis— sociation of an electrolyte a as a = A/AO where A is the measured equivalent conductance at a given concentration and A0 is the equivalent conductance at infinite dilution. At low concentrations and for a 1:1 electrolyte the Kd _ equilibrium AB 2: A+ + B led to the Ostwald dilution equation (72), 2 A C K = (1-17) d AO(AO-A) crfléi Ostwald dilution equation was rearranged by Kraus and Bray (73) to give, CA (1-18) A afld {... CE.“ o . l 010: c I age found to be u. electrolytes. 21 A0 and Kd can be obtained from the slope and intercept of a plot of % against C . The Ostwald dilution equation was found to be unsatisfactory for aqueous solutions of strong electrolytes. Kohlrausch (7H) found that for strong electrolytes the equiValent conductance approaches A0 according to the em- pirical equation, A = A - sol/2 (1-19) The square root law of Kohlrausch found its explanation in the Debye—Hfickel-Onsager (DHO) theory. Debye and Huckel obtained a good first approximation to the solution of the conductance problem from the electro- static theory of electrolytes. Debye showed that the eaguivalent conductance at low concentrations should be a liliear function of the square root of concentration. The next important step in the electrostatic theory of eslxectrolytes was taken by Onsager which resulted in the Chiisager limiting law for completely ionized electrolytes ('7ES,76). In the case of 1:1 electrolytes the Onsager lJirniting law is giVen by, A = A0 - (a*AO + 8*)01/2 (1-20) where in union r; is above theorgr a 22 where * _ 8.2 X 105 a " /2
3 * = 82 in which n is the viscosity of the solvent in poise. The above theory attributes the changes in the equivalent con- ductance with concentration to two long range effects of ion interactions known as the relaxation and electrophoretic effects. The relaxation term (%§ = -a*Cl/2) results from the perturbation of the applied field, X, by the asymmetry of the ionic atmosphere. The electrophoretic term (AAe = _B*Cl/2 ) takes into account the decrease in the velocity of an ion due to the counter flow of the solvent in the ionic atmosphere. According to this theory the coefficients a*, and 5* are independent of the model. A number of ex- tensions of the Onsager limiting law have been made (77— 85). The behavior of weak electrolytes was described by the Ostwald dilution equation while the limiting behavior of strong electrolytes could be predicted by the DHO theory. MacInnes and Shedlovsky (86) suggested that both the frac- tion of the solute that contributes to conductance and the mobility of the ions are concentration dependent. This led Fuoss and Kraus to propose that, "for electrolytes in general, equi increases, ‘00 tribute to 0 tion, and bees account of inc They introduce tance equatior electric const 0? conductance fluctance squat 4 an asSociated e A : AO~S(CQ)1/ Where LT] I 7? L77 f\) A) I l R) 5: 23 general, equivalent conductance decreases as concentration increases, both because the fraction of solute that con- tributes to conductance decreases with increasing concentra- tion, and because the mobility of the ions decreases on account of increasing effects of interionic forces" (87). They introduced the ion association constant to the conduc- tance equation to account for the deviations in low di- electric constant solvents, A = d(AO - SVCd) (l-2l) _ l-d KA ‘ 2 2 Ca Y+ where S = a*AO + 8*, and a is the degree of dissociation of the ion-pair. Fuoss and Onsager (81) extended the theory of conductance by adding higher order terms to the con- ductance equation. A revised Fuoss-Onsager equation for an associated electrolyte is (83), 1/2 A = AO—S(Cd) +ECdlog(Cd)+JlCd-KAA(Cd)yE-J2(Cd)3/2 (1-22) where E = ElAO — E2 E1 = 2.3026K2a2b2/2uc E2 = 2.3026K2abe*/I6Cl/2 K2 = nNe20/125 DkT b = e2/DkT and ‘5‘! thecoezzicier 24 l—q and KA = 2 2 Cd Yr _ bK/2 9.1’1Yi _ _ 1+Ka the coefficients J1 and J2 are functions of solvent prop— erties and can be expressed as a function of the size parameter a. There has been much debate on the necessity of introducing the J2(Co¢)3/2 term in the conductance equa— 'tion. This term has been partially calculated by different authors (83,84,88). The values for J2 and even for Jl obtained by different workers are not in agreement. Only accurate sets of data with internal precision better than 1% can be used with this method (89). The major limitation of the method for determining association constants is that the theory is valid only for symmetric electrolytes although it has been used for unsymmetric electrolytes with fair results (90). Equation (22) has been very successful in fitting conductance data of 1:1 electrolytes with 10 < KA < 1000 M.1 over a wide range of dielectric constants. For KA < IO M-1 almost horizontal plots of equivalent con— ductance versus 01/2 are obtained which makes the calcula— tions difficult. For KA > 1000 Mil the slope of the con— ductance plot becomes very high and insensitive to changes OfAd Also the computed sum, ECdlog(Cd)+-chd becomes neglected. Karl and conductance d necessarily n constant. higher-order" b paramebb_s, xvi. L«U t . WOPKEp 25 small with respect to KACayin, and therefore can be neglected. Karl and Dye (91) pointed out that a-good fit of the conductance data to the theoretical equation does not necessarily mean a good value for the ion-association constant. The reason is that both association and the higher—order terms involving ion size have the same concen- tration dependence in the first approximation and the separa— tion of the two effects depends upon second—order terms in the association (92). Karl and Dye showed that the terms which were dropped in the treatment of the electrophoretic effect are not small for those cases requiring the intro— duction of an association constant (91). In fact, simply retaining these terms in many cases yields surprisingly good fits of the data for reasonable and constant ion—size parameters, without requiring consideration of ion—pair formation (91). Since the publication of the original Fuoss-Onsager conduction equation (1957) various anomalies concerning the physical meaning of the distance parameter have appeared in the literature. The choice of ”a" used by Fuoss et a1. as the distance parameter was criticized by Prue and co— workers (93,9U). This criticism was supported by Justice and colleagues (95,96). A reanalysis of available con— ductance data by Justice (95) lead to a reconsideration of the physical meaning of the distance parameter in agreement Wit} the Blerrum ic the arbitrary distance (1: a Poisson equati values for the of q iS small. gested that th equation, 8, b of an ion-pail" Recently F separated ion-t tween unpaired the electrostat equation accorc eter R was deff beyond which cc I‘et’ion a g r : uncouples the F short range eff this model (aft t is prema c(inductance the present time is same value Of A Size Daramet er ; 26 agreement with the Bjerrum ion-pair concept. Even though the Bjerrum ion—pair model has been criticized because of the arbitrary cut—off represented by the choice of the distance q, a much more exact integration of the Boltzmann— Poisson equation (97) indicates that the range of possible values for the distance parameter around the Bjerrum value of q is small. To avoid any ambiguity, Justice (96) sug— gested that the distance parameter in the conductance equation, R, be called the "furthest distance of separation of an ion—pair". Recently Fuoss (39) proposed a model in which the solvent separated ion-pair is an intermediate transition state be— tween unpaired ions and a contact ion pair. He calculated the electrostatic and hydrodynamic terms of the conductance equation according to the above model. A distance param- eter R was defined as the distance from a reference ion beyond which continuum theory may be applied. Ions in the region a i r i B were considered as ion-pairs. This model uncouples the R parameter of the conductance equation from short range effects. The final conductance equation for this model (after some correction) is given by Fuoss (98). It is premature to assume that the difficulties of conductance theory are resolved. What can be said at the present time is that: (I) All equations yield nearly the same value of A0- (2) Those equations which use an ion— 3 size parameter similar to the Bjerrum distance are favored, . Fae-G‘- -y-s .‘Lu‘v' ‘ -:olvenr -¥\..Av’ “G 4.1 'D‘ . t . a :Ht. I t at: 27 (3) When association constants are to be compared, the same equation should be used to analyze the data, (A) when the association constant is large (>103) the choice of the conductance equation is not very important, and (5) Recent improvements in the statistical treatment of the problem are encouraging. (iii) Alkali Metal Nuclear Magnetic Resonance Spec- troscoEy (a) Introduction - Nuclear magnetic resonance spectroscopy is a powerful technique for the study of electrolytic solutions. Proton and 13C chemical shift and relaxation studies have yielded much information about ion—solvent, ion-ion and ion—ligand interactions in solutions. Alkali metal nuclear magnetic resonance spectroscopy can provide direct information about the interaction of alkali cations in solution. All alkali metals have natural isotopes with nonzero spin. Except for 7Li and 133Cs, line widths are dominated by quadrupolar broadening which means that deviations from a spherically symmetric environ— ment around the nucleus considerably broaden the NMR line. Lithium—7 and Cesium-133 show high sensitivity and the lines are generally narrow due to the small quadrupole moments of the nuclei. The chemical shift range of 7Li is about 10 ppm, while that of 1330s is several hundred ppm. Consequently resonance lines of dilute solutions (~ QQfiEfi; CH ANIV #:CEOE QCCQQCCO um I @UCTUCSQQ UT ANIEV .UOOUQQ SaUaZGCfi; OHCLZLCESG Cu ®>MQQH0EH HTLSQTZ .OQLE $52 .XOLQL< .KCLLL< HTOfiLuOQHE >Jfi>fi¢HmC®W -fiOHUZZ .nHQSH< -80 hhdctshflmuhns tHQeUHOJZ -H mofifldwru 29 .msso: wt: CH .mo93o oHQEmm EE OH cfi .pcoesppmcfl omlHpmHomH Handpmz .Uopm mzz .xosdd< .xosdd¢ Hmofisuooam mufi>fiufimcom .Hofiozz «mea< no mofiusodosm pmofiosz .H manna small relaxat make an appre hauser enhanc conclusion. mechanism is 30 For 133Cs similar studies (100) show that in spite of the small relaxation rate, dipole-dipole relaxation does not make an appreciable contribution. The recent nuclear Over— hauser enhancement study by Wehrli (101) confirms this conclusion. For the other alkali ions, the relaxation mechanism is mainly quadrupolar which is mediated by molecu— lar reorientation of the solvent. A large number of relaxation investigations have been 7 133 carried out for Li, 23Na, and Cs nuclei. According to the investigations of Hertz and coworkers (99,102,103), the 7Li quadrupolar relaxation rate for lithium chloride in aqueous solutions increases in the order I- < Br_ < Cl 23 << F-. The concentration dependence of Na relaxation in aqueous halide solutions (104) indicates that the relaxa— tion rate follows the sequence Cl’ < Br- 2 I- < F—. In the case of 133Cs the trend is Br” < I- 2 Cl- < F_ (102). In summary, for alkali halides (except fluorides) the concentration dependence of the relaxation rates is quite small and the sequence of 01—, Br- and I- ions bears no simple relationship to the ion size. Studies of the temperature dependence of the relaxation rate have also provided information on the potential barriers associated with the relaxation process (105,106). NMR relaxation studies of alkali ions were first re- ported by Jardetzky and Wertz (107). A 7Li quantitative relaxation study was performed by Craig and Richards (108) who studied 3 nethylforrnami been used in Eta+ (110,111: ions. These tightly 9&0?- ions in netha other solvent Since ion portant in no polar relaxer: vents should 3 vestigation o: relaxation net L. d 0 study 0: ion~a this subject ( equilibrium CO Uuuerstand' 31 . .+ . who studied L1 salts in methanol, formic acid, and di- methylformamide solutions. A variety of solvents has been used in the NMR relaxation studies of Li+ (106,109), + (112), Rb+ (110), and Cs+ (106,110) Na+ (110,111), K ions. These studies demonstrate the formation of a rather tightly packed first solvation shell for all the alkali ions in methanol, and for the lithium ion in a number of other solvents (113). Since ion-ion interactions are generally more im- portant in nonaqueous solvents than in water, the quadru- polar relaxation rates of these ions in nonaqueous sol- vents should provide very sensitive probes for the in- vestigation of such interactions. Although the quadrupolar relaxation method should be a very useful approach for the study of ion-association, the theoretical treatment of the ion-ion contribution to relaxation is very complicated. Therefore only a few investigations have been reported on this subject (106,109-111,11A). The reliable evaluation of equilibrium constants requires further development of the understanding of different contributions to ion-ion ef- fects (115). (0) Chemical Shift Studies - The shielding of a nucleus varies as a function of time due to ion-solvent, ion-ion, and ion-ligand interactions and therefore the observed chemical shift is an average over the different possible conff the chemical 5 both the probe '5 :3 O- (D F‘- 2:5 09 (I) (t s; C) distances. Th is expressed b developed a to 0v 32 possible configurations around the ion. To rationalize the chemical shift it is therefore essential to consider both the probability of various configurations around the ion being studied and the magnitude of the shielding con- tributions exerted by other species when situated at various distances. The total shielding that the nucleus receives is expressed by the screening constant 0. Ramsey (116) developed a general equation for the screening constant. Although his treatment can be applied quantitatively only to the simplest molecules, it provides an understanding of the origin of screening effects and a framework for further calculations. Saika and Slichter (117) have put the cal— culation of the screening constant on a practical basis by dividing it into three independent contributions, 0 = Gd + op + 0 (1—23) The diamagnetic contribution, 0d, arises from the local diamagnetic current in the molecule and is expressed by, k k|w0>l (1—2u) in which e and m are the charge and mass of the electron, c is the velocity of light, Ek is the angular momentum of the k'th ele k'th electro: magnetic terr . D .3 have cancticz 3 L A '- Oi Uhcse VTI‘JC’ protons is ‘4 usually the d approxinat e‘. V ....L) o = - D — #w-.a.~--eu.-ae ~~~ ~ 33 the k'th electron, and ‘rk is the radial distance of the k'th electron from the origin at the nucleus. The dia— magnetic term Gd depends only on the ground state electronic wave function To and is a function of the symmetry of the electronic distribution and the density of circulating electrons. Contributions to the shielding from the elec— trons of the other atoms are contained in GO. The effect and 0 on nuclei other than 3 of these two contributions, 0d protons if; usually minor. The paramagnetic term op is usually the dominant shielding term and can be expressed approximately as, A . A 3 p m C2 (wolk§,(£klk'/Fk)l¢o> (1‘25) Q II I l>|l-—’ where A is the average excitation energy. This term arises from the interaction of the ground state with excited states in the presence of a magnetic field. Several models (118,119) have been proposed in order to rationalize the large paramagnetic shifts of ionic crystals. It is not an easy problem to unambiguously choose the correct model, but quantitative estimates on the basis of different models have led most workers in the field (120—122) to prefer the overlap model proposed by Kondo and Yamashita (118). According to this theory the para- magnetic shift of cations and anions of alkali halides is s and p orbli Tons. '_1 .23 0Q ' .1 The GXtEI chemical Shi: Anions showed The order of ' 3“ due to the short—range repulsive forces between the outer s and p orbitals of the central ion and those of neighbor- ing ions. The extension of the theory of Kondo and Yamashita to chemical shifts in solutions is more complex due to random variation of the environment of the magnetic nucleus with time. In solutions an exact expression for the chemical shift would require knowledge of the radial distribution function of the solvent molecules and other ions around the central ion. Deverell and Richards (121) investigated the concentration dependence of the chemical shifts of aqueous alkali halides and nitrates by alkali metal NMR. Anions showed a definite pattern of shielding effects. The order of increasing shielding for all cations was I < Br < Cl < F‘ < H20 < N03 The magnitude of the shifts increased considerably with increasing atomic number of the cation. The authors concluded that the chemical shifts were caused by changes in interactions with both the solvent molecules and the counter ions. Typical plots of 133Cs chemical_shifts versus concentration for various cesium salts in waterenwegiven in Figure 1. Similar plots were obtained for other alkali salts. The concentration dependence of the 133Cs chemical shift of cesium salts has been examined by Halliday, Richards and Sharp (120). The authors assumed that the concentration dependence of the chemical shift is essentially determined by cation-anion .... FiEUl‘e l. 0631 COnQ ./ 133Cs Chemical Shift Csl/* in c/s at 7-350Mc/s f 400 .' / /)( Cs Br ‘ .//’ 300 i . / x _ ,s/ / Cs Cl / / ‘ 1‘ x/ 200-— // ‘/ / 1 {/i/ i/ /?f ,/{ .’// / F 1 / / / / M’ / / t ,l 100 I ‘ ./ {I /¢,t /‘/,¢ " i ‘/ 4/ f 0. "“h—:., CsNO3 L Concentration (moles/kg H20) 1 g g _1 1 2 3 I. Figure 1. Cesium—133 chemical shift as a function of concentration. Taken from Reference 121. hneractions accordin So 0') (I Cis t I$IIIIllIIIIIIIIIIIIIIIIIlIIIIIII----------_--==:____fl__“nb 36 interactions, and therefore, the concentration dependence of the chemical shift should follow the probability of cation-anion contact (HRS theory). By taking into account the influence of the ionic atmosphere on the probability of ion-ion contact through a Debye-Huckel type of treatment, the 133Cs chemical shift data were accurately represented from the lowest up to very high concentrations in H20, according to the following equation (120,123), I 2 ZaZ e expE-K(Re-a)] _ B _ I éobs — AC exp(— DkTRe 1+Ka ) — d + B (1—26) and A = 6(a) “—3'” (sfl — a3> Rm = a + RS 6(a) is a constant which depends only on the solvent properties, a and RS are the radii of the ion and the sol- vent molecule respectively, K is the inverse Debye length, C is the molar concentration, Za and Z8 are the valences éobs is the shift relative to the infinite dilution shift as zero, 6' of the ions, is an experimental shift referred to an arbitrary zero, B is the infinite dilution chemical shift, and Re (a < Re < Rm) has little effect on the con— centration dependence of the chemical shift. Later, Richards and coworkers (123) tested the validity of the HRS theory in low diele ..... no COPPGlatio l I , Shift ‘” 11“: Var U28)s of the 37 in low dielectric solvents and found it much less satis- factory. The experimental and calculated chemical shifts for CsBr in acetic acid-water mixtures are compared in Figure 2. The first systematic investigation of the applicability of alkali NMR in nonaqueous environments is due to Maciel 7 t 1. (12“) who determined the Li chemical shifts of LiBr, LiClOu in eleven nonaqueous solvents. Popov and co- workers (125,126) obtained the infinite dilution 7L1 chemi- cal shifts in a number of solvents by a systematic variable concentration study of various lithium salts. They found evidence for the formation of contact ion-pairs in tetra- hydrofuran, nitromethane, and tetramethylguanidine. Their results are in agreement with the data of Maciel at al. (12U) and of Akitt and Downs (127). In contrast to 23Na, no correlation was found between the limiting chemical shifts in various solvents and the Gutmann donor numbers (128)* of the solvents. The Gutmann donor number is an empirical scale of the donor ability of solvents. It is based on the enthalpy of formation of the 1:1 complex between a dilute solu- tion of a given solvent S and antimony pentachloride in 1,2-dichloroethane solution, 1,2 DCE s + Shel5 I S'SbClS (kca1.mo1“l) ’AHS-SbC15 = Gutmann donor number Figure 2. “’CB shift (obs) [Hz O 05 1.0 [CsBrJ/mol dm—a Cesium—l33 shifts against concentration for CsBr+HOAc-H20 systems (shifts referenced to infinite dilution shifts). o, 0.2 mole fraction HOAc, 0.8 mole fraction H20; 0, 0.4 mole fraction HOAC, 0.6 mole fraction H20, 0, 0.7 mole fraction HOAc, 0.3 mole fraction H20. Curves show the best least squares fit of data to Equation (26) by using solvent bulk dielectric constants: (a) D=57.2; (b) D=h2.5; (c) D=23.8. (Re=0.435 nm, a=0.365 nm, T=298K.) Note: curves a and b have the same origin but have been separated for clarity. Taken from Reference 123. The fir twelve nona Kidd (122). shifts in d the solvent gated vario DN- ‘vtass1u solutions q 39 The first comprehensive study of sodium iodide in twelve nonaqueous solvents was that of Bloor and Kidd (122). They found a correlation between chemical shifts in different solvents and the aqueous pKa values of the solvents. Popov and coworkers (125,129-132) investi- gated various sodium salts in a number of nonaqueous sol- vents. The experimental results of Bloor and Kidd (122) were confirmed. In addition, they indicated that the counter ion also plays an important role in determining the behavior 23 of Na chemical shifts. A linear relationship between 23 Gutmann's donor number and Na infinite dilution chemical shifts were obtained as indicated in Figure 3. This linear relationship was used to estimate the donor number of a number of solvents, which had. not been measured calorim- etrically, from the 23Na infinite dilution chemical shifts. Potassium-39 chemical shifts have been studied for KI solutions in ethylendiamine and KOH solutions in methanol (132). Recently Shih and Popov (133) presented a rather extensive report on the concentration and counterion de- pendence for a number of solvents. The concentration and counterion dependence of the 87Rb chemical shift of rubidium salts in nonaqueous sol- vents has not been studied extensively so far due to the broad lines. The solvent dependence of 1330s chemical shifts was briefly considered by Halliday et al. (120). A more F: We 3. Figure 3. U0 B \ P"\ III- \ t. \ n \ n!- \ \ "l ~ .‘2 \ n \ 9- \ 3' fl \ § -- . a 7-- \\ 1 ~ .. 5!- \ \ ot- \ \ 3'- \ \ 3b 7.\\ l)- \ Oh I! ..\ \ nII- \ -2. h\ ,3 L 1 1 g 1 1 1 1 1 1 O 4 I I! 16 N 36 29 33 36 no mm Sodium-23 chemical shifts Xi Gutmann donor numbers. Reference - aq. Na+ at infinite dilu- tion. 1. nitromethane, 2. acetonitrile, 3. acetone, A. ethyl acetate, 5. THE, 6. DMF, 7. DMSO, 8. pyridine, 9. hexamethylphosphoramide, 10. water. Taken from Reference 129. systematic aqueous soI The infinit nonaqueous dependent : contact ior exchange be on the NMR is obtained where 5f art and the ion cr [—1 V6 mole f tained from Where he mean act Debye‘HfiCke] obtained fI‘C Chemical 0f preferem Ml systematic approach to the 133Cs chemical shifts in non- aqueous solvents was reported by DeWitte gt al. (13“,135). The infinite dilution chemical shifts of alkali cations in nonaqueous solvent are summarized in Table 2. Concentration 133 dependent Cs chemical shifts were used to calculate contact ion-pair formation constants (135). Since the exchange between the solvated cation and ion-pair is fast on the NMR time scale, only one population averaged signal is obtained, 5 = ofxf + a X (l—27) obs ip ip where of and 6 are the chemical shifts of the free ion ip and the ion-pair respectively, and Xf and Xip are the rela- tive mole fractions of the two species which can be ob- tained from the equilibrium, Ki + - p + - M + X I M ' X (1-28) where l-a K = ip G2CYE The mean activity coefficient can be obtained from the Debye-HUckel equation. The values of Kip and 61p can be obtained from a nonlinear least squares iteration program. Chemical shift studies are also suited for investigations of preferential solvation phenomena in mixed solvent . a ... v . . _ "U. n i . 1|. 1n“ W; 0 AU L IJ I—IV ad AC a» .. a 3 n“ Cu AU Aw» .I. w .. [TV 10 L .J 3v ha ..- _ AL ma. ... u . ,1. fi H2 Table 2. Chemical Shifts/ppm at Infinite Dilution for 7 .+, 23Na+’ 39K+ and 133 L1 Solvents. Taken From Reference 115. + . . Cs Ions in Various Solvent Lia Na KC CsC Nitromethane —o.36' 45.6 —2l.lo —59.8 Acetonitrile —2.8O — 7 — 0.Hl +32.0 Dimethylsulphoxide —l.Ol — 0.11 + 7.77 +68.0 Propylene carbonate —0.6l - 9.U —ll.48 -35.2 Methanol —O.54 - 3.8 —l0.0S ~45.2 'Dimethylformamide 0.H5 — 5.0 - 2.77 — 0.5 Acetone 1.34 — 8.U —lO.H8 —26.8 Pyridine 2.5M 1.35 0.82 —3l aReference: Aqueous 4.0 M LiClOu. bReference: Aqueous 3.0 M NaClOu. cReference: Infinitely dilute aqueous solutions. systems. in the stu: 2: 13{ )3 Jfla preferentié quantity ir tion point' chemical st shifts of t asmmed the composition ionic solva of the two method <' I“ o p: that ion-ca; measuring t} 43 systems. Thus a variety of solvent mixtures have been used in the study of the alkali salts by 7Li (65,124,127,136, 23 87Rb (136) and 13303 (120,123,136, 137), Na (138—143), 137) NMR. In most cases only qualitative deductions about preferential solvation have been made. An illustrative quantity in this type of investigation is the ”isosolva— tion point" taken as the solvent composition where the chemical shift is at the midpoint between the chemical shifts of the ion in the two pure solvents (144). It is assumed that the isosolvation point corresponds to that composition of the binary mixture at which the inner cat— ionic solvation shell contains an equal number of molecules of the two solvents. One important point in using the above method to predict preferential solvation phenomena is that ion—pairing must be eliminated. This can be done by measuring the chemical shifts at different salt concentra— tions and, at each solvent composition extrapolating to infinite dilution. Popov and coworkers (143,145) have determined the isosolvation points of the sodium cation in a number of mixed solvents and found solvation to increase in the sequence, nitromethane << acetonitrile < pyridine < tetramethyl— urea 2 dimethyl sulfoxide z hexamethylphosphoramide A systematic approach to the preferential solvation has been made i simplifying alkali cati the equilit in which M Recently Be the chemica solvent mix bidentate a1 coordinatiol authors alsc of tetrahydl been made by Covington et a1. (136,137). By introducing simplifying assumptions the valuesof Kl/n were obtained for alkali cations in H202, H2O solvent mixtures, according to the equilibrium, n—X-lBX+1 + A (1—29) in which M is the alkali cation and A and B are solvents. Recently Delville gt al. (146) reported marked changes in the chemical shifts of NaClOu with the composition of binary solvent mixtures of tetrahydrofuran with unidentate and bidentate amines. Their results are consistent with tetra- coordination of the sodium cation by these solvents. The authors also followed the successive steps of displacement of tetrahydrofuran from sodium coordination by the amines. 2. Connie) Cryptar through the her of othe considered in addition ethers (fir; (first synt} DPOperties 1 tures of sor. compounds ar interesting CyCles for a in their use Of actiVe io The abil also had pro istry, which anion Salt b‘ the 4031; int: anionic, i e 45 2. Complexation of Alkali Cations by Crown Ethers and Cryptands A. Introduction Since Moore and Pressman (147) reported that the anti— biotic valinomycin induces the transport of potassium cations through the mitochondrial membrane by complexation, a num- ber of other naturally occurring macrocycles have also been considered as potential ioncarriersthrough membranes (148). In addition, a large number of synthetic macrocyclic crown ethers (first synthesized by Pederson (1-4)) and cryptands (first synthesized by Lehn and coworkers (4—7)) with similar properties have been prepared and investigated. The struc- tures of some naturally occurring antibiotics, crown compounds and cryptands are given in Figure 4. Particularly interesting is the strong and selective affinity by macro— cycles for alkali and alkaline earth cations which results in their use as models for carrier molecules in the study of active ion—transport phenomena in biological systems. The ability of crowns and cryptands to complex alkali cations, and their resistance to chemical reduction have also had profound effects on alkali metal solution chem— istry, which resulted in the isolation of the first sodium anion salt by Dye and coworkers in 1974 (149). Although the most interesting species in alkali metal solutions are anionic, i.e., solvated electrons and alkali metal anions, 46 O- ~hy¢uy Keno“... 3:: IL? Md Valinomycin redo (\o’} OF‘N’W E: “’3 c; 3 Ci :3 W0 \_/° 'vuv' l8-crown-6 lS-crown-S 1.10 diazo-lB-crown-6 c ° 7, W: W3 §§°W°W “Q32 cryptand-22l cryptand-222 CPYPtand—2ll r"’\ :3- 0 g 0 ("A p (— ‘ 7 wow Veg. OK’NVO-J cylindrical macrocylic spheroidal macrocyclic cryptand cryptand Figure 4. Naturally occurring and synthetic macrocycles it is the Cryptands studies of of simple the behavi Since : they have 3 Some of the the separat partitionir as catalyst tion reacti Crown e stoichiomet sium cation a2:l compl dihenzo~2u-. 30~crown-10 showed by IL plexes Witll in Which thE T .n this case dependent of ""4.- ” "ng‘v <__ 47 it is the complexation of alkali cations by crowns and cryptands which enhances the solubility. Consequently, studies of the thermodynamics and kinetics of complexation of simple salts with macrocycles can be used to predict the behavior of alkali metal solutions. Since macrocyclic ligands are expensive to synthesize, they have had limited commercial applications so far. Some of their possible commercial applications are (150): the separation of isotopes, separation of optical isomers, partitioning of radiactive streams. They can also be used as catalysts in electron transfer, and anionic polymeriza— tion reactions. Crown ethers are able to form complexes of different stoichiometries with alkali cations. For example the potas— sium cation can form a 1:1 (93%gfl) complex with l8-crown—6, a 2:1 complex with benzo—lS—cggwn—5, a 1:2 complex with dibenzo-24—crown—8, and a wraparound complex with dibenzo— 30—crown—1O as demonstrated in Figure 5 (151). Mei gt gt. showed by 13305 NMR that cryptand—222 forms two kinds of com— plexes with cesium salts (152). The inclusive complex is one in which the cation is inside the cavity of the cryptand. In this case the chemical shift of the complexed cation is in— dependent of the counterion and the solvent. The other form is an exclusive complex in which the cation is only par— tially inside the cavity so that the chemical shift still depends on the solvent and the anion (Figure 5). ’1. “We 5. 48 51' (\O/fi /\ O O 0:0 ° C @ 0) of? .0 OP'OVO K/Od L/ovo K+-18-crown-6 . (DB15-crown-5) 6‘0 0,8 Ofl 0 MO 0 . no gfia 0.169 . W . \ K§+'DB24-crown—8 + . K -DB30-crown-10 % Exclusive Cs+-C222 Figure 5. (a) Various stoichiometry of K+-crown ether complexes. (b) Exclusive Cs+-C222 complex. L. The de determinat of the reS' achieved b; are potent: spectroscoy such studie authors (15 general pri dynamics of in Section copy to stu ftrmation w The mos pounds is t: inDPEferen< tics of the sible fOP t} inflHence t} Cycle ligarx (1) fish 0‘ Dena" a1 Crow with those n1 49 The detection of the complexation reaction, and the determination of the thermodynamic and kinetic parameters of the resulting complex or complexes in solution can be achieved by a variety of physicochemical techniques. These are potentiometry, electrical conductance, calorimetry, spectroscopy, and relaxation techniques. The results of such studies have been reviewed extensively by several authors (153—159) and will not be repeated here. Instead general principles which govern the selectivity and thermo— dynamics of the complexation reactions will be discussed in Section (B). Then, the use of alkali metal NMR spectros- copy to study the thermodynamics and kinetics of complex formation will be described in Section (C). B. Selectivity of Complexation The most interesting characteristic of macrocyclic com— pounds is their ability to selectively bind certain cations in preference to others in solution. Various characteris- tics of the ligand, cation, anion, and solvent are respon- sible for this selectivity. The following parameters influence the selectivity and binding properties of macro— cycle ligands. (i) Relative Sizes of Cation and Ligand Cavity — In general crown compounds form the most stable complexes with those metal cations whose ionic radius best matches ¥ the radius tion (1,2: constants of cation cation rad Prewitt (l as the K-r stability alkaline e 50 the radius of the cavity formed by the ring upon complexa— tion (1,2). Figure 6 shows how the complex formation constants in the case of 18—crown—6 vary with the ratio of cation to cavity diameter (160). In this figure metal cation radii have been taken from the data of Shannon and Prewitt (161) and the cavity radii of complexes are taken as the X—ray crystallographic values (144). The maximum stability for complexes of l8—crown—6 with both alkali and alkaline earth cations occurs at a cation~to—cavity diameter ratio of unity. The increased stability of complexes of K+, and Ba2+ ions in aqueous solutions over those of other ions in the series is largely due to the enthalpy term (162) which corresponds to the greater electrostatic bond energy for those cations that fit the ligand cavity. For the smaller crown ether 15—crown—5, almost no cation selectivity has been seen since its cavity is too small even for the sodium cation. For benzo—l5—crown-5 it is difficult to distinguish the trends because of the formation of 2:1 complexes (Figure 7a). Larger crown ethers are not as selective as 18—crown— 6 because of the formation of complexes of variable stoi- chiometry. When stability peaks occur, as in the case of dibenzo-27—crown—9 and dibenzo—30—crown—10, the strongest Complexes are formed with K+ and Rb+ ions which are con— siderably smaller than the cavity of the two macrocycles (Figure 7b). In the case of cryptands, much better relationships FiEUPe 6 ‘ _.C1_:<.‘£D(/) 51 4.0~ 8.20 3.5- 3.04 I"°’W bgK 25s I? f) LOJ 2.0-4 Cavity Dianna-r Mom 1.5-< 1.0- c.» 0.5 If I 1 f 1 I I 0.7 1.0 1.3 Din. Mm/ Db. Cavity Figure 6. Selectivity of lB-crown-b: log K values for the re- action of 18-crown-6 with metal cations in H20 vs. ratio of cation diameter to 18-crown-6 cavity diameter. Taken from Reference 160. IogK Ck logK Cg C P(D.LL 52 5-1 ‘Z————21-Crown-7 in MoOH Dibenzo - 30-Crown—10 in MoOH (b) ‘.. Dibenzo- 21-Crown-7 in MeOH 24—Crown-8 in MeOH IogK 3 " Dibenzoeznt -Crown-8 in 70% MeOH 2 " Dibenzo-27-Crown-9 in 70% MeOH 1 l 1 ‘1 I 0 Cyclohexo-15-Crown-5 3 d (a) In MeOH O Dicyclohexo-12-Crown-4 . b9“ .\ inMoOH BmuoqumwnsinMeon 2 ' “(V o \\ o \ \‘ 1 - . ‘9‘ \i \15-Crown-5 in H,O I I r I l T u 100 130 150 | 180 Na‘ K‘ Rb' Cs’ Ionic Radius (pm) Figure 7. Selectivity of cyclic polyethers of various sizes: log K values for reaction of several crown ethers with alkali metal ions vs cation radius. (a) Crowns smaller than 18-crown-6. (b) Crowns larger than 18—crown—6. Data points labeled 0 indicate 2:1 complex forma— tion. Taken from Reference 155. eXiSt bet'l sizes of i strated iI bound to t tivity bec ably becat (ii) ethers, ev respect to of larger, than one c2 reduces the crown ethel results in ———i_ 7 - _ WM“; 53 exist between the stability of the complexes and the relative sizes of the cation and the cryptand cavity. As demon— strated in Figure 8, each alkali cation is preferentially bound to the cryptand with the proper size (162). Selec— tivity becomes less pronounced for larger cryptands, prob— ably because of ligand flexibility. (ii) Arrangement of Ligand Binding Sites — Crown ethers, even the small l5ecrown—5 are fairly flexible with respect to their oxygen donor groups in space. -The ability of larger, more flexible crown ethers to accommodate more than one cation or to wrap around a metal cation greatly reduces their selectivity. Cryptands are more rigid than crown ethers over a broader range of cavity sizes, which results in their higher selectivity. (iii) Typg and Charge of Cation — The binding of alkali and alkali earth cations by macrocycles can be considered to be electrostatic in nature. Many variations in coordination number and geometry are possible (Figure 5), and there are no real stereochemical requirements. All that is necessary is that the ligand provides an electron—rich en— vironment to replace the cationic solvation shell. Smaller cations, such as Li+ are more strongly solvated than the larger cations. Thus considerably more energy is required to replace the solvation shell of smaller cations. On logK g: *igure 8 7/‘nn1r 54 10.0-4 [2211b log K K‘ Rb‘ Cs‘ Metal Ion Radius (pm) Figure 8. Selectivity of cryptands: log K values for re— action of several cryptands with alkali metal cations vs cation radius. Data points: a — value reported <2.0; b — in 95% MeOH; c — in MeOH. Taken from Reference 162. C \ V J I 1‘ 7V 5 . d ' \ 4‘ I 9 'ani. cause of 1 6 other 00th becau . Y) L. ‘0; LEE 9994 ‘4...“ I 0 V th «LO e CO' V) 0 no t —: ' ___ 55 the other hand larger cations are not able to attract and ‘ organize the ligand molecules as well as smaller cations. These two effects cause the selectivity peaks for cations of 2+ to be generally higher than intermediate size, as K+, Ba for the other cations. Complexes of large alkaline earth cations often have higher formation constants than mono— valent alkali cations of similar size, whereas the op— posite is true when comparing small cations of different charges (163). Again this results from the competition between solvation and complexation. (iv) Type of Donor Atom - Substitution of sulfur or nitrogen for oxygen in the crown ether ring reduces the affinity of the ligand for alkali and alkaline earth cations, both because of the reduction of the cavity size, and be— cause of the lower donor abilities of S and N. However, this substitution increases the complexing ability of the ligands for transition metal cations of similar sizes, due to more covalent character in the binding (164). (V) Number of Donor Atoms — Little quantitative work has been done to investigate the result of varying the number of donor atoms in the ring without changing the size of the ring. Cram et al. showed that l8—crown—6 is a much better host for tert—butylammonium cation than is l8—crown- 5 with one less donor atom (165). Cryptand—222.produces much stronger complexes with alkali cations than its analog,cr chains ar (vi) and cowor benzene r of the ma: H: the ormai smaller t} to one of the Na+ cc ties of K+ benzene SU Stability 56 analog,cryptand—2208,in-which the oxygens in one of the chains are replaced by —CH2— groups (162). (vi) Substitution on the Macrocyclic Ring — Dietrich and coworkers (166) have shown that the addition of a benzene ring to crowns and cryptands alters the Selectivity of the macrocycles. For example, in methanol solutions, the formation constant of dibenzo-l8C6 with Ba2+ is smaller than with K+. The addition of one benzene ring to one of the bridges of C222 increases the stability of the Na+ complex in 95% methanol, but decreases the stabili— 2 ties of K+ and Ba+ complexes. The addition of another benzene substituent into a second bridge further decreases the 2 complexes (166). These results stability of the K+ and Ba+ may be explained in terms of ligand bulkiness, rigidity of the ligand and the electron withdrawing ability of the benzene ring. Substitution of cyclohexano groups has a less dramatic effect on the stability of the complexes and on the cation selectivity. Schori and Jagur—Grodzinski (167) have shown that in dimethylformamide solutions, the dinitro derivative of dibenzo—lB—crown—6 has a lower affinity for Na+ ion than the unsubstituted ligand. Conversely, substitution of -NH2 groups on the benzene rings results in a slight increase in the Na+ complex formation constant. These results are consistent with the electron—withdrawimgcharacter of --NO2 groups and electron-donating ability of —NH2 substuents. (Vii) reaction, the catio can produ constants the nature constants complex 57 (vii) Solvent and Anion Effects — In the complexation reaction, ligands must compete with solvent molecules for the cation in solution. Therefore, changes in the solvent can produce significant changes in the apparent formation constants of the complexes. The selectivity of the macro— cycles for certain cations over others may also change with the nature of the solvent. Frensdorff noted that formation constants of crown ethers with metal cations were three to four orders of magnitude higher in methanol than in water (16M). Kauffman et al. (168) found out that the enhance— ment of the stability of complexes in methanol over water is primarily of enthalpic origin which can be explained by the expenditure of less energy in the desolvation of the cation in a lower dielectric solvent. Cahen gt al. (169, 170) demonstrated that the dielectric constant is not the only solvent parameter that influences the stability of macrocyclic compounds. They showed that the solvating power of the solvent has a great influence on the stability of the complex. The degree of ion—pair formation of the complex affects the stability of the complex. In low dielectric media, and where the complex cation is exposed to the anion, com— plex formation competes with both the ion—pair formation of the salt and the complex and alters the stability of the complex. Smid and coworkers (171) found that in tetra- hydrofuran solutions, the complexes of substituted lB-crown- those wit the re‘SP predomine: undissocie of substit tetrahydrc The cc C221)+Pth .7 been measu (172). At two comple While the are 2-3 x 1 small d‘ '90s" .Lil complexed a unCOmplexec‘ VPOpy ‘ Complex 58 18—crown—6 polyethers with K+ are much less stable than those with Na+ ion, while in water and methanol solutions the reverse is true. In tetrahydrofuran, ion—pairs are predominent while in water and methanol we are dealing with undissociated cations. The order of stability of complexes of substituted 18—crown-6 polyethers with alkali cations in tetrahydrofuran is, Na+ >> K+ > Cs+ > Li+. The conductances of solutions of the cryptates (Na- C22l)+PhuB_ and (K C222)+PhuB— in tetrahydrofuran have been measured at various temperatures by Boileu et al. (172). At 20°C the ion-pair formation constants of the two complexes are 1.22 x 10M and 1.11 x 10“, respectively, while the ion—pair formation constants of uncomplexed salts are 2.3 x 10Ll and 1.1 x 10“. Considering the relatively small differences between ion—pair formation constants of complexed and uncomplexed salt, one may conclude that the uncomplexed salts form solvent separated ion—pairs (159). C. Thermodynamics of Complexation The thermodynamics of complexation of alkali cations by cryptands shows a remarkable range of enthalpic and entropic contributions. The contributions to enthalpy and entropy are discussed by Kauffman et al. (168). Ac— cording to these authors, the enthalpy of formation of the complex is influenced by: (i) placer moleCL outsic comple. (V) cation Changi] VGnt, eSDGC is eXDGCtec‘ thalpy Of c TL. b 18 expec independent 50f Refere Kauffma entropy 0f . (i) I the Catj 59 (i) The variation in the enthalpy due to the re- placement of the first solvation shell of the cation by the ligand. (ii) The change in the interaction with solvent molecules outside the complex as compared to those outside the first solvation shell of the cation. (iii) The change in inter-binding site (or intrasolva- tion shell) repulsions, which depends on the size of the solvent molecules and the degree of localization of the solvent dipole moments. (iv) The change in ligand solvation enthalpy upon complexation. (v) The steric deformation of the ligand by the cation. Changing the solvent from water to a poorer donor sol- vent, especially if the dielectric constant also decreases, is expected to yield negative contributions to the en- thalpy of complex formation from effects (i), (ii), (iv). It is expected that effect (v) will be relatively solvent independent. Effect (iii) is difficult to assess (Chapter 5 of Reference 157). Kauffman t 1. (168) attribute the changes of the entropy of cryptate complex formation to: (i) Entropy increase caused by the desolvation of the cation. by ori change (iv format amount freedo (V) StPUCtl The equilil by 0222 in (113), and Studied as Values of - Vents, reSp Of the 05+. to ‘14.l e. large negat wOrkers (17 ion With 1a; all c3368 t] tPOpy deSta] blentpopy t cmnbpmatior 60 (ii) Entropy increase caused by release of solvent bound to the ligand. (iii) Changes in ligand internal entropy caused by orientation, rigidification, and conformational changes. (iv) Decrease in translational entropy due to the formation of a single complex from two species, which amounts to -15 to —25 e.u., depending on the motional freedom remaining in the complex. (v) Decrease in solvent entropy caused by solvent structure formation about the large cryptated complex. The equilibrium constants for the complexation of Cs+ by C222 in acetone, propylene carbonate, dimethylformamide (143), and 95 wt% methanol-water mixture (162) have been studied as a function of temperature. The data yield AS° values of —32, —21, —19, and —23.7 e.u. for the above sol— vents, respectively. The value of A80 for the formation of the Cs+-18—crown—6 complex changesfrom -8.l e.u. in water to —14.l e.u. in methanol (173,17U). The origin of such large negative AS° value is not known. Popov and co— workers (175,176) have studied the complexation of the Cs+ ion with large crown ethers in nonaqueous solvents. In all cases the complexes were enthalpy stabilized but en— tropy destabilized. The authors assume that the decrease in entropy upon complexation is related to a change in the conformational entropy of the ligand, although it is not the only plexation of comple. action are and the me u) H? 61 the only factor governing the change in entropy of com- plexation. Many additional studies on the thermodynamics of complexation, as well as on the ligand—solvent inter— action are needed before the entropy of complexation and the macrocyclic effect (177) can be understood. D. The Use of Alkali Metal NMR to Study Thermodynamics and Kinetics of the Complexation Alkali metal NMR can provide unique information about stability, structure and dynamics as well as about kinetics of the complexation. Quantitative information on the stability of a complex can be obtained by fitting chemical shift, variable signal intensity or relaxation data to a model of equilibria. Upon complexation,the nucleus of the complexed cation usually resonates at a different frequency than that of the uncomplexed cation. When the exchange between complexed and solvated cationsis slow on the NMR time scale, two resonance lines are observed. The integrated intensities of the signals are proportional to the concen; trations of the two species, and thus the formation constant of the complex can be determined. However, this technique does not produce accurate results because of inaccuracies associated with the determination of the area under the peaks. When the exchange is fast on the NMR time scale, only one weigh complexed tion 01 ti fixed cat: formation the comple then the c in the cal \a in which 6 are the re Species. f while the ( Cationic SI F::::_________________________________________‘F5 62 one weighted—average signal occurs. If the free and complexed cations have different chemical shifts, the varia— tion of the chemical shift with mole ratio (M) at a cation fixed cation concentration can be used to evaluate the formation constant of the complex. If the cation and/or the complex exist in appreciable amounts as ion—pairs, then the contribution of these species should be included in the calculation of the observed chemical shift as i = Z X.6. (1—30) 1 in which éobs is the observed chemical shift, and the Xi's are the relative mole fractions of different cationic species. The latter are related to equilibrium constants, while the chemical shifts are characteristics of the various cationic species in solution. When ion—pair formation is important, the concentration dependence of the chemical shifts of the salt and the complexes will provide ion—pair formation constants. These data can then be used in mole ratio studies to obtain the formation constantcfi‘the complex. The variation of linewidth or relaxation times with the mole ratio could also be used to evaluate the complex formation constant. This method is particularly useful when one of the two sites has a much larger linewidth than the other (Chapter A of Reference (157)). .Alkali metal NMR spectroscopy also provides valuable informati a quadrup not have because 0 nucleus. function c thien f relatively line shape performed the quadru the linewi case only mate SOlut tiers (1114 alescence . the exChanE When the re DOSsible tc time. 63 information about the kinetics of complexation. When a quadrupolar nucleus is placed in an environment which does not have cubic or higher symmetry, the NMR signal broadens because of the asymmetry of the electric field at the nucleus. An investigation of the line broadening as a function of temperature yields kinetic parameters. When the cation exchange is slow, and the signals are relatively narrow, two resonance lines appear. A complete line shape analysis using modified Bloch equations can be performed to obtain kinetic parameters (178,179). When the quadrupole coupling constant of the nucleus is large, the linewidth of the complexed cation is broad. In this case only one resonance line is detectable, and an approxi- mate solution of the Bloch equations yields kinetic param- eters (114). The linewidth of the NMR signal at the co- alescence temperature also yields the approximate value of the exchange rate at this particular temperature (180). When the rate of the complexation is extremely slow, it is possible to study the kinetics of the complexation from the variation of the chemical shift or linewidth with time. CHAPTER II EXPERIMENTAL 614 The 1: times by f (181). Tl from a so] an ice-ace the crysta Shown in 1: hours at a nitrile. the meltin coldfinge. COOled nit] this Vacuur the cold fj bottle with 611133 kept (hiring Weig Was 39i1°c cryptand 22 cold finger 1. Purification of Materials A. Ligands The ligand 18-crown-6 (Parish) was purified several times by the formation of the complex with acetonitrile (181). The fine white crystalline adduct was precipitated from a solution of l8-crown-6 in acetonitrile by cooling in an ice-acetone bath. The solution was filtered rapidly, the crystals were transferred to the sublimation apparatus shown in Figure 9, and pumped under high vacuum for several hours at ambient temperature to remove weakly bound aceto- nitrile. Then the crystals were heated under vacuum through the melting point to about 60°C in an oil bath, while the cold finger of the sublimator was maintained at -50°C with cooled nitrogen gas. The l8-crown-6 crystals produced by this vacuum distillation were collected on the walls of the cold finger, and were transferred in a dry box to a bottle with a vacuum take off side arm. Purified 18-crown- 6 was kept under vacuum and in the dark at all times except during weighing. The melting point of the purified product was 39il°C in agreement with the literature value (181). Cryptand 222 (E. M. Laboratories, Inc.) was distilled under vacuum at 110°C and the crystals were collected on the cold finger and stored following the same procedure as in 65 66 To Vacuum Manifold Cold N2 Gas .\ Trap to protect Vacuum Manifold K) lmpure Figure 9. Cryptand sublimation apparatus. Taken from Reference CESE‘ ,o \d h L b .1 _ .1 Wm; AG owl. i ,(\ . C L r .Wl‘ . . w . C. 0‘ 11 03 d e e 0 .t d 13 t Arc .Q d e a S ab VJ P... hi 0 P b .l . . h. «l c x...” e in. .n . v... ad .5. e an u Lt hi 3 n S l .U. 0 ... . a; e to 1an S .l a (\ We w . o. . w; W... a hi . e i t m. S e -Tv d d .W. d _ a) no -. i e e a 3 . u e 71 S e ..i Va 6 . c .o .1 C. S «nu. n m...” nu P t t W U t 0 m Hr.“ HQ . Wm...” mm L ,U hu 0 C “k Ale 8 O Q . n 71‘ Vols .5 ..1 .1 t n . .1 a n. no . 0 Q h -1 m .nlu. 9 ac n? . 1.. LL. n1 ...uu w; 1. i ..V. la. 10 ill 8 «1‘ l: . 67 the case of 18—crown—6. The meltingpoint of the sublimed cryptand 222 was 68il°C (reported 68°C (182)). Liquid cryptand 211 was vacuum-distilled in an apparatus similar to that shown in Figure 9 except that a cup was attached to the cold finger (8). The cryptand was heated to about 65°C in an oil bath and the cold finger temperature maintained at —50°C in semidarkness. After distillation, the cold finger was warmed up to 30-35°C. The cryptand 211 then liquified and dripped into the cup. Upon cooling to room temperature, the ligand solidified as a waxy solid. The arms were broken and the cup was transferred to a bottle with a vacuum take—off side arm and stored in the same manner as other ligands. B. Solvents Methylamine (Matheson, anhydrous, 98%) was transferred from the gas cylinder to a glass bottle maintained under vacuum and containing Cal-l2 and a magnetic stirring bar. A pressure indicator was connected to the top of the glass bottle to check the H pressure build up as the solvent 2 was stirred over the CaHz. When the pressure indicator showed about 2 atmospheres, the solution was frozen with liquid nitrogen, and pumped out. After 2“ hours, the sol— vent was distilled into a bottle containing sodium—potassium alloy (1:3 w/w). Many freeze-pump-thaw cycles were per— formed until the liquid remained blue for several hours. If the so into anot The dry m glass bot Liqui< with Na-K over CaHQ ammonia. Methar day over m distilled. sulfoxide 68 If the solvent was not dry enough, it was transferred into another Na—K bottle and the procedure was repeated. The dry methylamine was then distilled into a heavy wall glass bottle for storage (183). Liquid ammonia (Matheson, anhydrous 99.99%) was dried with Na-K alloy in a similar way. The preliminary drying step over CaH2 was omitted because of the good quality of the ammonia. Methanol (Mallinckrodt or Fisher) was refluxed for a day over magnesium turnings and iodine and then fractionally distilled. Propylene carbonate (Aldrich) and dimethyl— sulfoxide (Fisher) were refluxed over calcium hydride under reduced pressure for 12 to 24 hours and then fractionally distilled. Methanol, propylene carbonate, and dimethyl— sulfoxide were further dried for 4-12 hours over freshly activated 4A molecular sieves. These solvents showed a water content of less than 100 ppm by Karl Fisher titra— tion. Deuterium oxide (KOR isotopes, 99.75 Atom %D) was used as received. The distilled water which was used for the calibration of the conductance cell was deionized by a "Milli—Q" water—purification system (Millipore Corporation). The specific conductance of the deionized water was found to be 9 x 10‘7 ohm—1 cm'l. The specific conductance of methylamine was 1.1 x 10'8 ohm"1 cm—l. u (\ Q a .C a .1 i 3: LI _ A... I. HA1. my. . ..H,H S “I“ Qv Ow . VIA e .a . AC ..nu mu \ e no a .c WV .1 mm n. .Q 7,. at. m. m nu DC a... n1 ..n1. .1- Flu AU .HA ..I. .nl. ... a S W; 9% n: w . O -H _ AC .nd 7: :1. said who .mu. m . ylborat S v I v L' Sl’llthesized Dhen S 73601“ Y? .3. 69 C. Salts Cesium iodide (Ventron Alpha Products), and cesium bromide (Polyresearch Corp.) were dried at 60°C under vacuum for several days. Cesium thiocyanate (Pfaltz and Bauer) was recrystallized from reagent grade methanol and dried under vacuum at 50°C for 2 days. Cesium tetraphenyl- borate was prepared by reacting sodium tetraphenylborate with cesium chloride in tetrahydrofuran solution. The resulting precipitate was washed thoroughly with distilled water and dried under vacuum at 70°C for 2 days (152). Cesium triiodide solutions were made by adding equimolar amounts of cesium iodide and iodine. Cesium nitrate and cesium perchlorate (Ventron Alpha Products) were dried at 120°C for several hours. Potassium chloride (Mallinckrodt) was recrystallized from Milli-Q water and dried in vacuum at 70°C, then ground and dried once more for two days at 200°C in the presence of P205. Rubidium bromide and ru- bidium iodide (Ventron Alpha Products) were dried at about 110°C for several days. Lithium perchlorate (Fisher) and lithium bromide (Matheson Coleman and Bell) were dried at 190°C for a few days. Tetraphenylphosphonium iodide (Ventron Alpha Products) was recrystallized from water and dried at 100°C. Tetraphenylarsonium tetraphenylborate was synthesized by mixing equal proportions of aqueous solutions of tetraphenylarsonium chloride (Aldrich) and sodium tetra- ‘phenylborate (Ventron Alpha Products), and dried at 100°C L2.) ‘..'v ..‘_ IIIII::::j——————————————————————————————————————————=‘-———————-————————————t===:: 70 for 2H hours. Tetraphenylphosphonium thiocyanate was synthesized by mixing equal proportions of aqueous solutions of sodium thiocyanate and tetraphenylphosphonium chloride, and dried at 50°C for 2H hours. 2. Glassware Cleaning The glassware for methylamine and ammonia solutions was first washed with an HF cleaner which consists of 33% HNO3, 5% HF, 2% acid—stable detergent, and 60% H20 by volume. After thoroughly rinsing with distilled water, the glassware was filled with aqua regia and kept under the hood overnight. The glassware was then rinsed thoroughly with deionized water and oven dried for several hours. Other glassware was soaked in sulfochromic cleaning solu- tion overnight, then rinsed with distilled water and oven dried. 3. NMR Techniques A. NMR Instruments All 133Cs and 7Li NMR data were obtained on a home— bUilt, single coil, pulsed spectrometer, at 1.A09 Tesla which employed a Varian DA—6O magnet and console, and was equipped with a wide band, variable temperature probe Capable of multinuclear operation (18A). A small external probe was placed 1.5 to 2.5 cm from the sample to serve ,- z.‘ .4 Width 01 .'_ 1 01°0th 1.0 71 as the lock. This probe contained water doped with a small amount of a paramagnetic species (185). The line— 1H lock signal was about A Hz. The Varian width of the proton lock circuitry was used to lock the magnetic field. A Nicolet 1080 computer which was coupled to a Diablo magnetic disk system was used to carry out the averaging of spectra and the Fourier-Transformation of the data. Cesium-133 and Lithium-7 NMR were performed at 7.87 MHz and 23.32 MHz, respectively. All 133CS NMR data were ob- tained with 5000 Hz sweepwidth and 8 K memory.. In all cases the linewidth at half-height was less than 2 Hz. The temperature was controlled within :0.5°C as measured with a calibrated thermocouple. The errors in chemical shifts were 30.15 ppm. In 7Li NMR measurements a sweep— width of 1000 Hz with 8 K memory was used, and the chemical shifts obtained were accurate to within i0.l ppm. A11 87Rb NMR and one set of 133Cs NMR (for a comparison with DA-6O results) measurements were performed on a Brucker WH-180 superconducting multinuclear spectrometer which consists of a superconducting solenoid, a Nicolet 1180 computer disk system, a quadrature detection system, and a temperature control unit. Most of the spectra were obtained with a sweepwidth of 5000 Hz and used 8 K of 87Rb chemical shifts depend memory size. The errors in the on the linewidth and the signal to noise (S/N) ratio of the :resonance line. The broader the line, the smaller is the 72 S/N ratio, and less well—defined is the maximum of the peak. B. Data Handling Cesium—133, lithium-7, and rubidium—87 chemical shifts were first referred to; a 0.7 M aqueous solution of CsBr, a 4.0 M aqueous solution of LiClOu, and a 1.0 M solution of RbBr in D20. The chemical shifts were then corrected to refer to infinitely dilute aqueous solutions. The values of such correctionsforl33Cs, and 87Rb reference solution are, +9.89 ppm, and +5.89 ppm, respectively. The reference solution for 133Cs measurements was sealed in. a 5 mm NMR tube which was coaxially sealed into a 10 mm tube. The space between the two tubes was evacuated and vacuum—sealed. In this way, the ambient temperature shift of the reference could be obtained even when the probe was cold (186). Two kinds of reference samples were used for 7Li measurements. Both of them contained 4.0 M aqueous LiClOu solutions but one of them was placed in a 10 mm NMR tube and the other in a 5 mm tube which was coaxially sealed into a 10 mm tube. The chemical shift of aqueous LiClOu is concentra- tion independent; but because of the homogeneity differences, the chemical shiftscfi‘these two reference samples are not 0.23 ppm). All 7Li chemical shift the same (a a 10mm— 5mm = data were corrected to the infinite dilution with respect netic sus: reference Transform (1}) 73 to the reference sample in the 10 mm NMR tube. Chemical shifts are corrected for the differences in bulk diamag— netic susceptibility of the solvent (nonaqueous) and the reference (aqueous) (l87a) as appropriate for Fourier Transform NMR (l87b), 2 sam 1e ref 5 — + 3710“) p - xv ereme) (1—31a) corr _ 6obs 41 ( sample _ reference) 3 XV Xv 6corr = éobs (1'31b) where Xsample Xreference v , u are the volume susceptibility (31b) of the sample and reference solutions, respectively, 6corr and éobs are the corrected and observed chemical shifts, respectively. Equation (l—31a) applies to the DA—6O instrument where the applied magnetic field is transverse to the long axis of the cylindrical sample tube, while Equation (1-31b) applies to the WH—l80 spectrometer where the polarizing magnetic field is along the long axis of the sample tube. The values of volume susceptibility of the solvents and the corresponding corrections on each instrument are given in Table 3. The contribution of the salt to the magnetic susceptibility of the solution was ignored since all measurements were done at low concen- trations. .WDQGESCHQWCH Omwfllzz UCQ OWI L0 COflUOTLLOU EQHHfiQHHQanmwumw OfiQQEMwWEWufiQ -m @HQMRH 74 mx m>+<>\m> + >x m> + <>\<> n >x < oLprHE "509% oopmHSonQm ems.o hopes se.o- mmm.o+ Aev woe.o omzoxteaseaseeesxoa w:.o| :m.o+ mom.o Aomzmv oUHKOMHSmfikcooEHQ mm.o: wH.o+ :mm.o, Aomv ouwcoohmo oCoH%Q0hm mm.OI m:.o+ mam.o AmOoEV Hocmflpoz fim.o+ moa.ou mm.o masoEE< s:.o| mmm.o+ wow.o AooHHIV ocflEmamnpoz AEQQV AEQQV OH x zpflafloflpdoomsm pco>aom owfllmz go om| xasm .mpgoszspmcH owalmz ocm ow|flom mSOHsm> mo SOHpoohhoo hpflafipflpgoomsm OHrommEaHQ .m oHomE lllIllIlIIlIIIIIIIIIIIIIIIIIII-IlI--------- The i were obta equations gram (188 Two k were used Solutions wall this: solvents 1 Of 0.5 mm 0°C a8 Sh< measuremel 75 The ion-association and complex formation constants were obtained by fitting the NMR data with the appropriate equations using the nonlinear least squares KINFIT4 pro- gram (188) on a CDC-6500 or a CDC-7501 computer. C. NMR Sample Preparation Two kinds of 10 mm 0.D. precision NMR tubes (Wilmad) were used for measurements on the DA-60 NMR instrument. Solutions in liquid ammonia were prepared in tubes with wall thickness of 1.0 mm; those in methylamine and other solvents were prepared in tubes with the wall thickness of 0.5 mm. These two kinds of tubes were calibrated at 0°C as shown in Figure 10. The samplesftm°Brucker WH—l80 measurements were prepared in l5-mm spinning NMR tubes (Wilmad) and coaxially mounted in 20 mm tubes which con- tained D20 as the lock. All NMR samples in methylamine and liquid ammonia were prepared under high vacuum (<10.5 torr) in extended NMR tubes of the type shown in Figure 11. In the case of salt solutions, solutions used for the mole ratio studies, and solutions used for the concentration study of the 2:1 1806 (“T Cs ligand were placed into the NMR tube. The Kontes valve ) complexes, weighed amounts of the salt and/or the was then closed and the tube was connected to the vacuum manifold with 9 mm Fisher-porter ”Solv—Seal" joints. The NMR tube was flamed and pumped out for an hour. Then the Au.— 6.0 ' 3.0 - L n 0 0. 9.. I SEE» @E3.0\/ figure 10 76 a) E _ .. g4o- - 2 _ - § l 1 l l I l I 6.0 8.0 l0.0 Height (cm) I l l I I 1 3.0 *- N O l Volume (ml) '6 I 0 [0 2.0 3.0 40 5.0 6.0 Height (cm) jFigure 10. Calibration curve for 10 mm NMR tubes at 0°C. (a) wall thickness = 0.5 mm, (b) wall thick— ness = 1.0 mm. Figure 11. 77 T0 VACUUM g Figure 11. Extended NMR tube for high vacuum. solvent :msimw l—cJ i“: ’..' |__.l thorough the sclu 78 solvent was distilled very slowly into the NMR tube, which was immersed in an isopropanol—dry ice bath at -78°C, and filled to the mark. The walls of the tube were washed thoroughly by the solvent to bring all the reagents into the salution. The solution was then frozen completely in liquid nitrogen, pumped out, and vacuum sealed. The frozen' solution was thawed in an isoprOpanol-dry ice bath. The height of the solution in the tube was measured carefully at 0°C, and finally the volume of the solution was obtained from the calibration curve (Figure 10). In the case of the concentration studies of the 1:1 and 2:1 complexes, stock so- lutions were made in methanol. Known volumes of the solution were transferred to the NMR tube, frozen, and pumped out. Methanol was then removed by distillation under high vacuum, and the residue was pumped out for 1 hour, before transfer 'of the solvent into the NMR tube. Salt solutions in sol- vents other than methylamine and liquid ammonia were pre— pared from stock solutions by using volumetric flasks. The volumetric flask was weighed before and after the addition of the ligand, then the flask was filled to the mark with the stock solution of the salt. The sample for the kinetics study of complexation of lithium bromide by cryptand-211 was prepared in the apparatus shown in Figure 12. Weighed amounts of lithium bromide and the ligand were placed in the NMR tube and the side-arm respectively via the exten- sion tubes. The extension tubes were sealed and the Figure 12 . 79 TO VACUUM _ +- Figure 12. Extended NMR tube for the kinetic experiment. apparatu into the the vacul side-arm both the (‘1‘ the sal apparatus salt and tions we: ratus was The time data was data were data vau 0V er a tW ' Com RGSiS' HZ with a by ThOmDS< minor mm the Cell I was shunts 80 apparatus was connected to vacuum. Methylamine was distilled into the NMR tube and the apparatus was disconnected from the vacuum. Some of the solvent was distilled into the side-arm and the apparatus was shaken at -50°C to dissolve both the salt and the ligand. Care was taken to ensure the salt and the ligand solutions were not mixed. The apparatus was kept at -50°C overnight to make sure that the salt and the ligand were completely dissolved. The solu- tions were mixed at -50°C in the NMR tube, and the appa- e ratus was placed in the precooled (-50°C) probe immediately. The time lag between mixing and the acquisition of the first data was about 15 seconds. After a few pulses (1-5) the data were stored on the disk for later processing. The data acquisition and storage process was repeated 98 times over a two hour period. 4. Conductance Method A. Conductance Equipment Resistances were measured at 400, 600, 1 K, 2l Pm wflflEwfihflpmz CH HmO .HO P.HHQD HQOHEBC) D) aD|.|SIL|)1 Ill and shc in the magneti interac solvent phenylb illustr sible t! The 8011 was fine; Solutior lil Com; chemica; practice fOre> Ct COUCentr mole rat Table 6 iHCPeasi at Vario CGSium 0 than Wit CeSium t: independ. the ChEm; aPe giVeI 89 and shown in Figure 1“. At all temperatures an increase in the concentration of cesium iodide results in a para- magnetic (downfield) shift indicating that the cesium cation interacts more strongly with the iodide ion than with the solvent. The data for a similar study with cesium tetra- phenylborate in methylamine are given in Table 5 and are illustrated in Figure 15. Since the solubility of cesium tetraphenylborate in methylamine is low, it was not pos- sible to study this system at temperatures below -3°C. The solubility of cesium tetraphenylborate in methylamine was increased by the addition of lB-crown-6 to the salt solutions. Since the formation constant of the resulting lzl complex is larger than 10“ (Chapter IV), a plot of the chemical shift versus (l8-crown-6)/(Cs+) mole ratio is practically linear in the range 0 < mole ratio < 1. There- fore, the chemical shift of the uncomplexed Salt at each concentration could be obtained by extrapolating to zero mole ratio. The results of such studies are shown in Table 6 and Figure 16. A diamagnetic shift occurs upon increasing the concentration of cesium tetraphenylborate at various temperatures (Figure 15) which means that the cesium cation interaction with the solvent is stronger than with the tetraphenylborate anion. In the case of cesium thiocyanate, the chemical shift is concentration independent in the concentration range studied. However, the chemical shift is temperature dependent. The results are given in Table 7 and illustrated in Figure 17. In ) hobsflppm "4 X "6* "8- .\ 120- |22~ |24~ 126- 6‘ )3. .28 . ’2) 63% 4‘ ’3’5 wok Fi‘a’ul‘e lu ppm) __’___________ ___.._L. 90 .gure l“. “4 ..Il6 -||8 l20 |22 |24 l26 l28 I30 Concentration and temperature dependence of the Cs chemical shift of cesium iodide in methyl- amine. E / E), u. . v? a l 5 / \ 0C 9. MU- «3 7i ii 0 5 7 0 5 9 mi a v1 C 0 II. _,{ 7 1!. 2 7a 7. 6 .l. 9 0 0 WI. dd 33 4|. m e . .Hv O 0 .Il qli l 2 2 «3 3 C). PW; r0 7! O m .l. C O «No AU 0 0 AU 0 flu O 0 AU filo 0 0 TL 6 e b n 0 0 Av flu fl filo AU AU AU 0 O 0 AU 0 0 h h a O . . . . . . . - . . - C at T nb flu flu ml; 0 Wu flu O 0 0 O O 0 AU 0 dd 2“ 91 133 Table 5. Concentration Dependence of the Cs Chemical Shift of CsBPhu in Methylamine at Various Tem— peratures. éobs (ppm) Temperature, 0C Sonc. (M) 25.0 13.2 5.8 -2.9 0.00048 17.72 24.70 27.96 31.37 0.00072 13.30 19.35 22.69 25.71 0.00074 l3.07,l3.07 22.53 26.02 0.00113 9.19 14.62 17.80 20.82 0.001275 7.72 13.38 16.25 19.27 0.00171 6.56 14.70 17.26 0.00210 4.85 9.97 12.84 15.55 0.00265 2.83 7.88 10.36 12.68 0.003175 1.75 6.48 8.65 10.90 0.00390 0.04 6.79 9.27 3.00505* 2.0:0.5 3.8t0.5 1.00509' —2.44 3.00612 —4.07,-3.91 2.45 ).00727* —5.0i0.5 -O.95i0.5 l.OiO.5 ).0103* —7.u:o.5 -3.4 :0.5 1.65:0.5 3Chemical shifts were obtained from the extrapolation of the mole ratio curves to 0.0 mole ratio. .1gure 15 «EQQV anon Sobs (ppm ) Figure 15. 92 3O -. 1 l 1 l 1 1 1 1 l 1 l 0.0 0.2 0.4 0.6 0.8 l0 l2 [Cs’] x lo"- (M) Concentration and temperature dependence of the 133Cs chemical shift of cesium tetraphenylborate in methylamine. U able 4‘ e w I n AC 1.) .l w I! AU a U 4|. #0 O O b . Q - AU DIV 0 a 93 Table 6. Variation of the 133Cs Chemical Shift with the Mole Ratio (18C6)/(CsBPhu) in Methylamine Solu— tions at Different (Cs+) Concentrations and Various Temperatures. Gobs (ppm) (Cs+) Mole Ratio Temperature, °C (M) + (1806)/(CS ) 25.0 13.2 5.8 0.00505 0.0a 2.0e0.5 3.8:0.5 0.404 8.03 10.12 8.11 0.519 9.50 11.75 0.648 11.75 13.92 0.760 13.61 15 55 0.00727 0.0a _5.0:0.5 —0.95:0.5 1.0:0.5 0.396 2.76 6.48 8.57 as 0.456 4.00 7.49 9.81 0.668 8.11 11.52 13.85 0.822 10.28 13.69 15.78 0.0103 0.0a -7.4:0.5 —3.4:0.5 —1.65:0.5 0.209 —3.06 0.74 2.60 0.495 3.07 6.79. 8.73 0.644 6.01 9.43 11.52 0.857 9.89 13.54 15.71 aObtained by the extrapolation to 0.0 mole ratio. 4.1 I t igure 1( 94 a) [Cs’] . 00051 M b) [cg - 00072 M v I 1 l at of g 5 - E 5 - 5 § 10 - ‘° IO - aw c are 102- c '5 P- 5“ '5 '- 5'. C 1 1 . 1 J . 1 1 1 00 04 08 00 04 on [mm/[cc] [necsmcr] c) [(23] - 00103 M ' I ' I -5- o— (I Figure 16. Cesium-133 chemical shift versus (18-crown- 6)/(CsBPh ) mole ratio at different concen- trations of CsBPhu and various temperatures. 0.001. 314 u]. 0 0.01803 L4_4, rrrrr‘———————————————-——————————__‘=‘=“‘_“’_____:flllillv 95 Table 7. Concentration Dependence of the 133Cs Chemical Shift of CsSCN in Methylamine at Various Tem— peratures. aobs (ppm) m Temperature, °C conc. (M) 25.00 9.4 6.0 -2.5 —10.3 —16.2 0.00094 60.68 64.33 65.26 67.51 68.82 69.76 0.00148 60 68 64.17 65.34 67.04 68.82 69.91 0.00428 60.76 64.25 65.34 67.12 69.06 69.76 0 0054 60.65 —————————— 67.15 68.89 ————— 0.00823 60.53 64.48 65.34 67.27 68.75 69.99 3.01314 60.76 64.56 65.41 67 04 68.59 69.99 3.01803 61.07 64.40 65.41 67.58 68.82 69.99 a a. a. K ..r. A ELQQV «new 3 . .1gure l ...a 8.8. 96 58. I f I I I I l I T 60- 5°C 7’5 5‘5 WA 0 25.0°C 627 A 6 t- a g- 4 30 T o 9.4oc 0- 3 "’ 4 v .. tr 6.0°C 3 66- ' O 00 " O o o J 0 O ‘2-5OC 68‘ -wr o 4 ° -|03°c 7O -°L 1* c # -l6.2°c 72 1 #14 L 1 4 1 l 1 0.0 0.4 0.8 IE 1.6 2.0 [w] x 102 (M) ‘igure 17. Concentration dependence of the 133Cs chemical shift of cesium thiocyanate in methylamine at various temperatures. addit101 shift 0: S.6°C wz shown in ical shi nitrate of these of the 5 results bilities To c data for 19. T} should, fore, tr is ObVic occurs 6 cumulati Conseque Solution Chemical bOPate a thiOCYah determin The P980 The made hum 97 addition, the concentration dependence of the 1330s chemical shift of cesium triiodide in methylamine at 25.0°C and 5.6°C was studied briefly. The data are given in Table 8 and shown in Figure 18. The concentration dependence of the chem— ical shift of cesium bromide, cesium perchlorate, and cesium nitrate could not be studied due to the low solubilities of these salts in methylamine. Therefore the chemical shifts of the saturated solutions were measured at 25.0°C. The results are given in Table 9 along with approximate solu— bilities of the salts in methylamine. To compare the results, the concentration dependence data for various salts at 25.0°C are illustrated in Figure 19. The chemical shift of the free solvated cesium cation should, of course, be independent of the counterion; there- fore, the curves should converge at infinite dilution. It is obvious from Figure 19 that substantial ion—association accurs even at the lowest concentration studied, where ac- :umulation times of 8—10 hours per sample was required. Ionsequently, it was not feasible to study even more dilute 1olutions. From the concentration dependence of the 133Cs .hemical shift of cesium iodide and cesium tetraphenyl- crate and the requirement that these salts and cesium hiocyanate have a common intercept, it was possible to etermine the ion—pair formation constants of these salts. he results will be presented in Section C. The long extrapolation of the data to infinite dilution 1de numerical calculations very difficult, so that attempts 98 Table 8. Concentration Dependence of the 133 4444V______________—___________—==fi‘__________:illll’ Cs Chemical Shift of CsI3 in Methylamine at 25.0° and 5.6°C. 6(ppm) Temperature, °C Conc. (M) 25.0 5.6 0.00131 123.88 123.57 0.00348 125.59 125.36 0.01177 128.46 127.37 0.01991 131.64 130.93 |22 l24 l26 |28 7801.. (ppm) l30 I32 1. 56°C: 250°C 1 l 1 l l '3‘004 0? 08 1.2 .gure 18. T6 20 [or] x 102 (M) Concentration dependence of the 13305 chemical shift of cesium triiodide in methylamine at 25.0 and 5.6°C. 100 Table 9. Cesium—133 Chemical Shifts of Saturated Solu- tions of CsBr, C8010“, CsNO3 in Methylamine at 25.0°C. Salt Conc. (M) . dobs(ppm) CSBP <0.001 103.80 103.87 CscloLI 50.0007 49.52 CSNO3 <0.0007 46.72 46.34 CsBph4 ‘ .1 40 ‘_ sat. CsNO3 - , g " sat. CsClO4 _ 3 60 -- : : : fl.- CsSCN " egg" — _ 80 - — '00 2 $01. CsBr ' 14K) 1 1 1 I 1 1 1 1 I 1 1 0.0 04 0.8 12 LG 20 2.4 ire 19. Concentration dependence of the [Cs*]x102 (M) 133 Cs chemical shifts of some cesium salts in methylamine at 25.0°C. 102 were made to obtain the chemical shift of the solvated cesium cation experimentally. Although such attempts.failed to give the chemical shift of the free cation directly, the data contain some information about ion-association which will be discussed in Section C. Cesium thiocyanate data do not appear to give much in- formation because of the concentration independent chemical shift. In the case of cesium thiocyanate three possibilities exist: (1) Cesium thiocyanate is not ion-paired even at the highest concentration. From the behavior of the other salts in this low dielectric constant solvent this possibility is highly unlikely. The linewidths of the signals are larger for this salt than for the other salts which is also an in- dication of ion-ion interactions; (ii) It is possible that cesium thiocyanate is com- pletely ion-paired even at very low concentrations. If this is true then the chemical shift might suddenly change at lower concentrations which are beyond the detectability limit of the method. The direction of the change is not known even though in all the solvents which were studied by Popov and coworkers (129-135) a diamagnetic shift has been observed for alkali thiocyanates; and (iii) It is probable that the chemical shifts of the free solvated cesium cation and cesium thiocyanate ion pairs are the same. This would not be surprising since both 103 :hiocyanate and methylamine are nitrogen donors. Actually, ;his assumption formed the basis for our choice of the :hemical shift for the free solvated cesium cation as rill be discussed later in this chapter. B. Ion Association of Cesium Salts in Methylamine Various models were used in an attempt to analyze the :oncentration dependence of the 133Cs chemical shifts of :esium iodide and cesium tetraphenylborate in methylamine Lolutions. Among them, a model which involves the forma— 1ion of ion—pairs and two kinds of triple—ions is the most 11ausible and will be discussed in detail in Section ii—b. lhort reviews of two other models will be given in sections and ii-a for the sake of completeness. (1) Simple Ion—Pair Model — We first tried to fit the hemical shift data for cesium iodide and cesium tetra— henylborate with a simple model in which only ion—pairs orm. The following equations were fit to the data by sing the weighted nonlinear least squares program KINFIT: K. 1p Cs + X Z CS-X (3-1a) (CS.X) = l—CX (3-lb) 1p (Cs)(X)vi d2Cy2 i" 104 6 1/2 Y1 = exp _(__£1121§fl£19_£921_____I7?_) (3—1c) ‘——_‘—__————1 3/2 50.298100) (DT) [1 + (DT)1/2 6 X + obs = CsaCs XCS°X605.X = 06CS + (1— )GCs.X (3—10) in which Cs and X are the solvated cation and solvated anion, respectively; Cs-X is the ion—pair (charges are omitted for simplicity), the terms in parenthesis in Equation (3—lb) are the molar concentrations of various species, Kip is the Lon-pair association constant, a is the degree of dissocia— ;ion of the ion—pair, Y: is the mean activity coefficient of :he salt in solution, C is the analytical concentration of 1he salt, D is the dielectric constant of methylamine, T is he temperature in , 8 is the distance of the closest ap— roach (8 = 5.3 h)*, sobs is the measured chemical shift, Cs and 6Cs-X are the chemical shifts of the free solvated ation and ion—paired cation, respectively, and X03 and 05.x are the relative mole fractions of the free and ion— aired cation respectively. Values of the dielectric onstant at different temperatures were obtained from the quation in Reference 194, D = DO - 0.09195 (3—2) The distance of the closest approach was chosen as the av— ’age of (rCS+ + r1-) and ¢Cs+ + PI' + rs) in which rCs+ = ‘67 8 and r _ = 2.20 8 are the crystal ionic radii of the ésium cation and the iodide anion respectively, and rs = 7 is the Vander Waals radius of methylamine molecule. 1wever, the exact choice of "a” does not affect the results .gnificantly. 105 where DO — 11.3 is the dielectric constant of methylamine at 0°C and t is the temperature in oC. Cesium iodide and cesium tetraphenylborate data at 25.0°C were fit simul— taneously by equations (3-1) with five adjustable param— eters. These are: two ion—pair association constants, two chemical shifts of the ion—pairs, and the chemical shift of the free cesium cation in methylamine (see Ap— pendix 1A for details). The results of this type of curve fitting are illustrated in Figure 20, where the experimental and calculated chemical shifts are compared. The adjust— able parameters are given in Table 10. The simple ion—pair model requires a sharp change in the chemical shift at low concentrations with a flat portion at high concentrations. Even the "best fit" of the data is very poor at all tempera- tures. A closer look at the chemical shift—concentration plots (Figures 14 and 15) shows the existence of rapid Changes with concentration at low concentrations and slower continuing changes at higher concentrations. This indi— cates that other interactions are important in the solu— tion. Table 10 shows that none of the adjustable param— eters are well—determined by this model. The average standard deviation of the residuals is about 1.0 ppm vhich is much higher than the experimental errors in the zhemical shift determination. Therefore it is clear that 1 simple ion-pair model cannot describe the behavior of lesium salts in methylamine, and that other effects are 106 1 I r l . I I l I I . T I CsBPh4 4O - E 60 '- CL - £- 3 80- o m - IOO~ 120K. ' _ C51 140 - l I l I l l l I 1 I 1 I 1 0.4 0.8 l2 |.6 2.0 2.4 [C§]x l02(M) Figure 20. Concentration dependence of 133Cs chemical shiftscfiszI and CsBPhh in methylamine at 25.0°C. o experimental points, x calculated from the simple ion pair model. 107 Table 10. Calculated Thermodynamic Parameters of Ion- Association of Cesium Salts in Methylamine at 25.0°C According to a Simple Ion—Fair Model. CsI CsBPhu .9 .. 6Cs+ ppm ~0115 K' M-l (“i )x10“ (14:1)x10Ll 1p 5- ppm 130:: _33i25 1p 108 also responsible for the chemical shift changes as a func- tion of concentration. It must, however, be realized that most of the changes in chemical shift with concentration can be accounted on the basis of ion-pair formation so that the remaining information needed to improve the model is relatively minor. It would seem that the logical next step would be to invoke triple-ion formation. (ii) Formation of Triple—Ions - According to theory (89), the maximum concentration, Cmax’ for which triple- ion formation of a univalent electrolyte in a solvent of dielectric constant D may be neglected is given by: Cmax z 1.19 x lO-lu(DT)3. In methylamine, even at the lowest temperature, and concentration which we studied, the formation of triple-ions cannot be ignored according to this equation. Therefore, we tried to analyze our data according to models in which triple-ions form. (a) One Type of Triple-Ion - In this model it was assumed that only ion pairs and anionic triple—ions will form in the solution. Even though the formation of cationic triple—ions cannot be neglected, this model was used as a starting point and is worth describing briefly. The equilibria and subroutine equations required for the analysis of multiple data sets according to this model are given in Appendix 18. 109 To analyze the data completely (neglecting the tempera- ture dependence of the chemical shifts), it is necessary to evaluate at least seven parameters for each salt. Four of these are ion—pair and triple-ion formation constants together with the enthalpies of the formation of the two species. The other three are the chemical shifts of the free solvated cation, the ion-pair, and the triple-ion. Although we found out that our data could be fit well by this model, the determination of all seven parameters is not possible. In general, it becomes apparent in all model- fitting that those parameters which are highly correlated with each other, and/or are characteristics of minor species in the solution cannot be determined simultaneously. For example, high correlation coefficients exist between the formation constant of the anionic ion-triple, K and its ta chemical shift. Since the anionic triple-ion, X.Cs.X, is a minor species in the solution, we cannot determine both the chemical shift and its formation constant simul— taneously. Because of the existence of only small relative amounts of Cs+ in solution, even at the lowest concentra- tions, and the consequent long extrapolation of the chemical shift to infinite dilution, the accurate determination of the chemical shift of the solvated cation was not possible. If we try to adjust 6Cs+’ high marginal standard deviations in the other parameters result. Although we would prefer to determine the chemical shift of the free cation by data 110 fitting, it was necessary to fix it at a reasonable value. In order to maketflmeproblem tractable, we assumed that the chemical shift of the ion-pair of cesium thiocyanate and that of the free cesium cation are equal (see section iii for more explanation of this choice). In addition, the association constant of the triple—ion was calculated from the Fuoss equation (24) and used as a constant (more eXplanation is given in section iii). With these assump- tions, the data at various temperaturesnmawsfit well by this model and the average standard deviation of the re- siduals was reduced to 0.45 ppm. The calculated thermo- dynamic parameters obtained from 11 data sets are given in Table 11. Even though all parameters, except the chemi- cal shift of the triple-ion, are well-determined by this model, we feel that the formation of the cationic triple- ion cannot be ignored. Therefore,we refined our model by introducing the formation of the cationic triple—ion in the equilibria. (b) Two Types of Triple-Ions - The most reasonable model for the analysis of the NMR chemical shifts of cesium salts in methylamine is a model in which the ion-pair Cs+'X-, and both kinds of triple-ions (anionic and cationic) exist in the solution. The equilibria involved in this model are, 111 Table 11. Calculated Thermodynamic Parameters for Ion- Association of Cesium Salts in Methylamine at 25.0°C According to the Ion-Pair and Anionic Triple-Ion Model. Assumptions: 1. 6C5 = 60.73 ppm from CsSCN data at 25.0°C. 2. Kta = 32.5 M_l from the Fuoss equation at 25°C. CsI CsBPhu Kip M‘l (4.3+1.0)x105 (1.510.51x10“ Ahgp kcal. mole-l (2.0:0.3) (4.710.5) écs-X ppm 121.7i0.4 -2l:5 SX-CS°X ppm 784ill9 -337:58 112 Kip Cs + X z Cs-X C + CS°X°CS (3-3) where X-Cs-X and Cs-X'Cs are anionic and cationic triple- ions (charges are omitted for simplicity), and Kta and KtC are formation constants for the anionic and cationic triple- ions, respectively. The equilibrium constants, mass balance equations, charge balance equation, and activity coefficient equation were used to solve for the concentration of the different species in solution. The numerical calculation was based on an iterative method with convergence on the major species (Cs+-X‘) in the solution. The detailed algebra and the subroutine EQN for the KINFIT program are given in Appendix 10. The calculated concentrations of the various species were then used in the chemical shift equation, 2 X. = 1 (3—4) 1 113 where Xi's and 51's are the mole fractions and chemical shifts of the species containing the cesium cation. A complete solution to the above equations for each salt at various temperatures would require the adjustment of three association constants, three enthalpies of formation (for the ion-pair and two kinds of triple-ions), and four chemical shifts even if the temperature dependence of the chemical shifts is ignored. Since the chemical shift of the free cesium cation is a common parameter for both salts, 19 parameters would have to be adjusted for a complete. solution to the problem. As mentioned in Section ii, the high correlation among some of the parameters would clearly prevent solution of the complete problem. Assumptions must therefore be made in order to obtain any quantitative in- formation about the system. As we shall see, certain parameters are very insensitive to these assumptions and can therefore be well-determined. The final set of assump- tions was: 1. The chemical shift of the free solvated cation was taken as the chemical shift of cesium thiocyanate in methyl- amine. The concentration independent chemical shift of cesium thiocyanate suggests that the chemical shift of the free solvated cation and the ion-paired cation are the same. It seems that the chemical shift of the triple- ions and the ion-pair are also the same, since there is 114 no concentration dependence. The chemical shift of a 0.02M cesium thiocyanate solution in methylamine which is saturated with tetraphenylphosphonium thiocyanate is 60.84 ppm at 25.0°C Which is equal to that of cesium thio- cyanate in methylamine. The addition of tetraphenylphos- phonium thiocyanate to the solution should increase triple-ion formation. Only if the chemical shift of the ion-pair and the anionic triple—ion are the same would we expect the same chemical shift upon addition of tetraphenyl- arsonium thiocyanate. This experiment suggests that the chemical shift of the triple-ion and the ion-pair are equal, so that the chemical shift of the ion-pair and the free cesium cation would also be expected to be the same. A similar experiment was carried out for a cesium iodide solution in methylamine. The chemical shift of a 0.024M solution of cesium iodide in methylamine which is also saturated with tetraphenylphosphonium iodide is 129.38 ppm at room temperature while the chemical shift of pure cesium iodide at this concentration is 127.06 ppm. A down- field shift (2.2 ppm) as a result of the addition of tetra- phenylphosphonium iodide indicates an increase in the forma— tion of the triple—ion. The addition of tetraphenylarsonium tetraphenylborate to a 0.0032M cesium tetraphenylborate in methylamine, results in a 0.38 ppm upfield shift at 25°C, also indicating the formation of more triple-ions. As mentioned above, the chemical shift of the free solvated cation, 5 could not be obtained from the data Cs’ 115 since a very long extrapolation is required (Figure 19) and the concentration of the free cation is very small (less than 10% at the lowest concentration used). However, the choice of 6Cs+ is critical, since ion—pair association constants are affected significantly by a change in the chemical shift of the free cation. For example, if we change 60s from 80 ppm to 40 ppm, the ion—pair formation constant of cesium iodide decreases by a factor of 4 while the ion—pair formation constant of cesium tetraphenylborate increases by the same factor (compare Tables 12—14). The residual surface is not very dependent upon the choice of one of these values, provided the other parameters are adjustable,and both choices of 008 give nearly equally good fits to the data. In summary, our choice of the chemical shift of the free solvated cesium cation is (003+)t = 60.73 - 0.224 (t — 25.0) (3-5) where (0CS+)tis the chemical shift of the free cesium cation at a given temperature. The temperature dependence was ob- tained from a plot of the chemical shift of cesium thio— cyanate versus temperature which yielded a straight line with a slope of -0.224 as shown in Figure 21. 2. Equal probabilities for the formation of two kinds of triple—ions. The determination of the formation 116 60 70- 72 l l l I ~20 -10 0 1o '20 Figure 21. cyanate solutions versus temperature. Cesium—133 chemical shifts of cesium thio— 117 Table 12. Thermodynamic Parameters for Ion-Association of Cesium Salts in Methylamine at 25.0°C; Ob— tained with the Assumptions that Ion-Pairs and Two Kinds of Triple Ions are Present, and That 6Cs = 60.73 ppm at 25.0°C. Assumptions: 6Cs = 60.73 ppm (from CsSCN data ' at 25.0°C). _ _ —1 Kta — Ktc — 32.5 M (from Fuoss equation at 25.0°C). 6Cs.x.0s = 60s.x Adjustable Parameters at 25.0°C CsI CsBPhu K. M‘1 (2.54i0.31)x105 (1.41:0.30)x10” 1p Ahgp kcal. mole‘l (3.86:0.28) (4.71i0.40) = . i . 1 -1 .1:4.2 SCS-X-CS 6CS-X ppm 124 19 0 3 3 4 8:24 — 2 8:42 GX'CS‘X ppm 5 9 IIIIIIIIZZ::44r———————————————————————————__________lll .11, 118 Table 13. Thermodynamic Parameters for Ion—Association of Cesium Salts in Methylamine at 25.0°C; Obtained with the Assumptions that Ion—Pairs and Two Kinds of Triple Ions are Present, and That 6C5 = 40 ppm. Assumptions: 605 = 40 ppm. Kta = KtC = 32.5 M"1 (from Fuoss equation at 25.0°C). 6Cs.x.Cs = 6Cs.X Adjustable Parameters at 25.0°C CsI CsBPhu Kip M—l (4.3:0.6)x105 (7.6:0.7)x103 AHEp _ kcal. mole"l (3.40:0.3) (4.66:0.3) écs-x-Cs=5cs.x ppm 124.05i0.33 -9.8i2.3 GX'Cs-X ppm 545i4o —336i21 119 Table 14. Thermodynamic Parameters for Ion—Association of Cesium Salts in Methylamine at 25.0°C; Obtained With the Assumptions that Ion—Pairs and Two Kinds of Triple Ions are Present, and That 6Cs = 80 ppm. Assumptions: 0C5 = 80 ppm. Kta = Kto = 32.5 M‘1 (from Fuoss equation at 25.0°C). 60s.X.Cs = 50s.x Adjustable Parameters at 25.0°C 1 CsI CSBPhu -1 5 4 Kip M (1.10:0.13)x10 (2.97io.67)xlo AHEp kcal. mole‘l <4.8510.27> (5.38:0.57) 6Cs-X-Cs=SCs-X ppm 125.07iO.3 —8.4i3.4 _. + 6X-Cs-X ppm 359:11 391—77 120 constants of the triple—ions .isnot possible along with the determination of the ion—pair formation constants, since these two constants are highly correlated to each other and only a small fraction of the cesium cation is in the form of the ion—triple (less than 10% at the highest concentra— tion). The values of the ion—pair formation constant, enthalpy of the formation, and chemical shift of the ion— pair do not change significantly with a change in the forma— tion constant of the triple-ion. Therefore the assumption of equal values of the two triple—ion formation constants is acceptable even though the degree of the solvation of the cation and the anion are different. 3. Fixed value for the formation constant of tgiplg:ign§. Since triple ions are minor species and the ion—pair parameters do not change much with a Change in the formation constant of the triple ion, we can fix the formation constants of the triple—ions at ”reasonable" values with the understanding that the resulting chemical shift parameters for the triple—ions will depend upon this choice. Two procedures were used: (a) The formation constants of the triple—ions at different temperatures were calculated from the Fuoss equa— tion 2WNa3 3 = W I933) “-8) 121 Values of I(b3) for integral values of b are tabulated (24) so that values of I(b3) for nonintegral valueswereobtained from a plot of log 1(b3) versus b. (b) Independent of the model which we used, the value of the formation constants of the ion—pairs at 25.0°C ob— tained from our NMR data for cesium iodide is almost 5 times larger than the Fuoss value, and almost 5 times smaller than the Fuoss value for cesium tetraphenylborate in methylamine. If we assume that the ratio of the ion—pair and triple—ion formation constants calculated from the Fuoss equation is the same as the corresponding ratio obtained from NMR data, we can write, K = A x K. (3—6) where A = (Kt/Kip)Fuoss' This assumption suggests that if any effect is missing in the calculation of the ion—pair formation constant according to the Fuoss theory, it is also missing in the calculation of the triple-ion formation constant. Again, it must be emphasized that some such assumption is forced upon us by the inability of the chemical shift data to define all the unknown parameters. The analysis of data according to either assumption 0 or b yields the same thermodynamic parameters for ion— pair formation (compare Tables 12 and 15). If we adjust 122 Thermodynamic Parameters for Ion—Association of Table 15. Cesium Salts in Methylamine at 25.0°C; Obtained with the Assumptions that Ion—Pairs and Two Kinds of Triple Ions are Present, and That Kt/Ki is a Known Constant. p Assumptions: 60s = 60.73 ppm (from CsSCN data at 25.0°C). = _ -3 A Kt/Kip — 6.8 x 10 (from Fuoss equations at 25.0°C). CSCs.X.Cs = 6Cs.X Adjustable Parameters at 25.0°C CsI CsBPhu —1 5 4 Kip M (2.68:0.19)X10 (l.l6:0.l5)x10 AHEp kcal. mole_l (3.75i0-ll) (4.0910.l5) 0 = 0 Cs x-Cs Cs.X ppm 124.05iO.20 —17.8i3.2 5 ppm 299il.2 —906i51 X'Cs-X 123 "A" for the best fit, the standard deviations become too high but the best-fit value of this parameter is equal to the corresponding value calculated from the Fuoss equations within the standard deviation (Table 16). Even if we use a value for the triple-ion formation constant which is 3 times larger than the Fuoss value, no change in the ion— pair formation parameters occurs. This confirms the idea that the choice of the triple-ion formation constant is not important as long as its value is within a reasonable range. 4. Chemical shifts of anionic and cationic triple- ions. The high correlation between the chemical shifts of the two kinds of triple-ions prevents their determination simultaneously. For example, since they are formed in equal concentrations, the chemical shift of one could go down an arbitrary amount provided that it is balanced by a corresponding increase in the chemical shift of the other triple-ion. Therefore we were once again forced to make some assumptions about the values of their chemical shifts. It is important to point out that the choice of the chemical shifts of triple ions does not influence the ion- pair formation parameters significantly. Essentially the same values (within standard deviations) for the ion-pair formation parameters result regardless of the choice of the chemical shifts of the triple-ions. The reason is that the 124 Table 16. Thermodynamic Parameters for Ion—Association of Cesium Salts in Methylamine at 25.0°C; Obtained with the Assumptions that Ion—Pairs and Two Kinds of Triple—Ions are Present, and that Kt/Ki is an Adjustable Parameter. p Assumptions: 5C8 = 60.73 ppm (from CsSCN data at 25.0°C). A = Kt/Kip (Adjustable) 50s.x.0s = CSCs.x (6Cs.X>t = (6Cs.X)25.O + b(t‘25'0) Adjustable Parameters at 25.0°C CsI CsBPhu —l 5 4 Kip M (2.85i1.0)x10 (1.05i0.35)X10 Apr kcal. mole_l (3.65il.4) (2.56i3-2) acs-X-Cs=6CS-X ppm l23.73:1.19 —19.9i8.8 b ppm-deg‘l —0.0067:0.0029 -0.25iO.53 6X-Cs-X ppm 283ill ~608i273 A 0.0011110.00055 125 ion-pair formation parameters are almost completely characterized by the low concentration region of the chemi- cal shift—concentration plots, and in this region less than 1% of the cesium cation is present in the form of triple- ions. Most workers in the field agree that alkali metal NMR chemical shifts are only sensitive to nearest neighbor interactions. Reasonable assumptions about the chemical shifts of the triple-ions can be made by considering the nearest neighbor effect. a. If we assume that the degree of overlap of the outer p-orbitals of the cation and anion are the same in the ion-pair and the cationic ion-triplet, Cs.X.Cs, then the chemical shift of the two species should be equal. With this assumption the chemical shift of the cationic species could be fixed at the ion-pair chemical shift and 5 the chemical shift of the anionic triple-ion could be adjusted. b. If the degree of overlap is not the same for ion—pair and ion-triplets, then we can assume that a change in the degree of overlap results in a change in the chemical shift. Thus the chemical shift of the cationic triple- ion can be expressed as CS°X°CS ‘ 60s x 126 where A is the chemical shift characteristic of the dif- ference in the degree of overlap between the cation and anion in the ion-pair and ion-triple. Since the cesium cation is surrounded by two anions in the anionic triple- ion, the chemical shift of the anionic triple ion can be written as GX-Cs-X = 2((SCS.X - 5C3) + 6C5 - A (3_7b) assuming that the same effect-causes corresponding changes in the chemical shifts of both triple—ions with respect to the chemical shift of the ion-pair. We used both assumptions a and b in the analysis of the NMR data. Both assumptions lead to the same values for the ion-pair formation parameters, again showing that the choice of the chemical shift of triple—ions is not critical (com- pare Tables 12 and 17). 5. Temperature dependence of the chemical shifts. Since the degree of interaction between cation, anion, and the solvent is temperature dependent, the chemical shift of various species should also be temperature dependent. Un- fortunately there is no theory for the temperature depen- dence of the chemical shift. The CsSCN data show a linear dependence of the chemical shift on temperature (Figure 21). In general, it can be assumed that the chemical shifts of 127 n—Association of Table 17. Thermodynamic Parameters for To Cesium Salts in Methylamine at 25.0°C; Obtained with the Assumptions that Ion—Pairs and Two Kinds of Triple Ions are Present, with Equi— partition of the Chemical Shifts. Assumptions: 50st: 60.73 ppm (from CsSCN data at 25.0°C). Kta = Kto = 32.5 M‘1 (from Fuoss equation at 25.0°C). 50s.x.Cs = 5Cs.x ‘ A 5x.0s.x = 2(50s.x ' 50s) + Cs ‘ A Adjustable Parameters at 25.0°C CsI CsBPhu —1 5 4 Kip M (2.49i0.28)x10 (1.49i0.31)x10 AHEp kcal. mole‘l (3.89i0.26) (4.89:0.41) écs-X ppm 124.28i0.29 —11.8i3.8 A ppm -l33i11 114125 7__—4. ,1. r ..raerQ - _. 1,, 128 X-Cs-X Table 18. Thermodynamic Parameters for Ion-Association of Cesium Salts in Methylamine at 25.0°C; Obtained with the Assumptions that Ion-Pairs and Two Kinds of Triple Ions are Present, and that SOS X is Temperature Dependent. ° Assumptions: 6C = 60.73 ppm (from CsSCN data 5 at 25.0°C). Kta = KtC = 32.5 M.1 (from Fuoss equation at 25.0°C). 6Cs.X.Cs = 5Cs.X (SCS.X)t = (50s.x)25.0° + b(t'25'0) Adjustable Parameters at 25.000 081 CsBPhu -1 5 4 K10 M (2.50:0.22)x10 (l.21:0.2l)x10 AHEP kcal. mole”l (3.15:0.21) (3.3lt0.79) = 4. i . - . t . GCs-Xocs 6Cs-X 12 47 0 23 16 1 3 7 b 0.020i0.002 -O.l6i0.09 0 469il9 -280i27 129 various species is linearly dependent on temperature. The chemical shifts of triple ions are not well determined by any of the models. Therefore it is obvious that the intro— duction of temperature dependence of the triple-ion chemi— cal shifts is not warranted. A comparison of Tables 14 and 18 also indicates that the ion-pair formation parameters are the same whether the chemical shifts of the ion-pairs are temperature dependent or not. Lest the reader feel that any collection of parameters can fit the model, it should be emphasized that the model chosen must fit not only the concentration dependence of the chemical shifts of CsI and CsBPhu at a given temperature but at EL; temperatures studied. While the data do not per- mit unambiguous choice of all possible parameters, the model used is a reasonable one - theory predicts that ion—pairs and triple-ions should form in this solvent. We expect different chemical shifts for free ions, ion-pairs, and triple-ions and reasonable choices of the corresponding chemical shifts suffice to describe the data very well. C. Complementary Experiments Since the determination of the chemical shift of the free solvated cation in methylamine requires a long ex- trapolation of the data, we tried various ways to determine it experimentally. All such attempts failed, but the data provide some information about ion—association of cesium 130 salts in methylamine. Therefore, the results will be given in this section. i. Cesium Tetraphenylborate in 90% v/v Methylamine in _____________________________________________ Dimethylsulfoxide Solutions — It was thought that if we could extrapolate the chemical shift-concentration data of cesium salts in mixed solvents, where one of the solvents is methyl— amine, then the chemical shifts of the free cesium cation in various solvent compositions could be obtained. Extrapo— lation of these chemical shifts to 100% methylamine would then give the chemical shift of the free cesium cation in methylamine. Exploring this idea, we measured the 13303 chemical shift of cesium tetraphenylborate in 90% v/v methyl- amine in dimethylsulfoxide solutions as a function of the salt concentration. The data at 25.09C are given in Table 19 and shown in Figure 22. It had been found previously that cesium tetraphenylborate does not form ion-pairs in DMSO and that the chemical shift of the cesium cation in this solvent is 68 ppm (135). From the data of Table 19, it is obvious that cesium tetraphenylborate remains highly associated in this mixed solvent, probably because of the high percentage of methylamine. The chemical shifts at all concentrations are about 40 ppm downfield from the cor— responding values in pure methylamine solutions. Even though the solvent contains only 10% DMSO, the chemical shift is determined largely by the chemical shift of the cesium 131 Table 19. Concentration Dependence of the 13305 Chemical ggiggcof CSBPhh in 90% V/V Methylamine-DMSO at [08+] (M) obs (ppm) 0.00056 57.04 0.00075 54.17 0.00105 50.83 0.00135 47.58 0.00174 47.11 0.00221 45.25 0.00255 44.48 0.00352 42.30 0.00428 41.30 0.00527 39.44 0.00615 39-59 0.00674 38.70 132 s -. 0.50- m .. .0 0 00 60 r- r . .4 l I 1 .I l 0.0 0.2 0.4 0.6 0.8 [Cs *] x IOZ (M) Figure 22. Concentration dependence of the 13303 chemical shift of cesium tetraphenylborate in 90% v/v methylamine in dimethylsulfoxide at 25.0°C. 133 cation in DMSO indicating preferential solvation by DMSO. It is clear that the extrapolation of the data to infinite dilution cannot be easily done. Therefore, this experiment did not help us to obtain the chemical shift of the cesium cation in methylamine solutions. (ii) Mixtures of Cesium Iodide and Cesium Thiocyanate in Methylamine - Early in the study of cesium salts in methylamine, when we had examined only the concentration dependence of the chemical shift of cesium iodide and cesium thiocyanate, it seemed plausible that cesium iodide forms contact ion—pairs in methylamine while cesium thiocyanate does not (see Figure 19). To test this hypothesis, mixtures of these two salts at a constant total concentration of the cesium cation but various mole ratios of the two salts in methylamine were prepared and the chemical shifts measured. The results are given in Table 20 and are compared with the results for the pure salts at 25.0°C in Figure 23. Extrapo- lation of the cesium iodide data to high concentrations gives a chemical shift of 127 ppm or larger. If the limit— ing shift were due to ion-pairing only, a plot of(6 -127)/ obs (GCSSCN-127)versus the mole fraction of iodide should be a straight line with unit negative slope, provided the pres- ence of CsSCN shifted the CsI ion-pairing equilibrium to completion. The fact that the plot is curved (Figure 24a) shows that this simple interpretation is invalid. If,on 134 Table 20. 133Cs Chemical Shifts of Different Mole Ratios of (I‘)/(SCN‘) +(I') in Mixtures of CsI and CsSCN in Methylamine at Various Temperatures; (Cs+) = 0.0054 n. _ éobs (ppm) (_I ) _ Temperature, 0C (SCN )+(I ) 25.0 —2.5 -10.3 0.00 60.65 67.15 68.85 0.26 77.67 81.85 83.02 0.52 92.40 95.04 95.73 0.76 107.75 108.61 ' 108.92 1.00 122.10 122.33 123.26 135 _ [1-]/ [scw-]+ [1'] ‘Eg (SC)-o.-——-o—o-———o—-—4L ’rOChfiscwi g ‘000 § ' a 026 ,0 80- b A 0.52 ICO- _ ' 0.76 '20“ .11“: lg CSI ”4() 1 1 .1 1 I J .1 1 .1 I 1 1 00 0.4 0.8 l2 LG 20 24 [05*] x 102 (M) Figure 23. A comparison of 13305 chemical shiftscfi‘pure cesium salts and the mixture of cesium salts in methylamine at 25.0°C. 80b5— 8CsI Sosscw- 8CsI Figure 24. .0 (I) 1 ~ (0) - 0.6r / 4 L _ (b) 4 1' - 1- _ 0.2 - - 136 I I I I I 1 O 0.2 0.4 0.6 0.8 1.0 XI‘ Plot of (éobs—SCslj/(acsSCN _6CsI) versus mole fraction of iodide. (a) 6CSI = 127.0, (b) écsl = chemical shift of CST at the con- centration of CsI in the mixture. 137 the other hand, cesium thiocyanate is more strongly ion- paired than cesium iodide, a plot of gomho o .QOHpoELoa somloadflap mcfluooawo: Ugo ANH oHDoBV whopososog pHoQIQOH Sosa oopoHSoHoo x .m.o u IHX e.o_xo._ q A2108 m-o_xo._ .QOHpoAQCoocoo go comp 0.0. 5.. _ To. x O._ n.o_xo._ _ Om wk wk. Vb Nu. ON mm mm vm Nm .sm opsmflm (011111me 143 formation (the solution to the equilibrium equations become very difficult when all possible types of triple ions are considered). The differences in the calculated and experi— mental chemical shifts can be accounted for by the forma— tion of triple ions. For example at a total concentration of 0.003 M, and the mole ratio (I-)/(BPhu-)=l, according to our model, almost 3.7% and 0.7% of the cesium cation is in the form of iodide and tetraphenylborate triple-anions respectively. Since the chemical shifts of ion-pairs and anionic triple—ions are different, the existence of triple- ions results in a chemical shift. The value of this chemi- cal shift is about +1.8 ppm which accounts for the dif- ferences in the calculated and experimental chemical shifts. D. Conclusion Cesium iodide and cesium tetraphenylborate are highly associated in methylamine. According to theory triple— ions form in methylamine as well as ion—pairs. A model in which ion-pairs and two kinds of ion-triples are formed, fits well to our NMR data. To obtain the ion—association param- eters of these systems, it is necessary to make certain assump— tions. Among the various assumptions mentioned in this chap— ter only the one concerned with the choice of the chemical shift of the free solvated cesium cation is critical. The assumption that the chemical shift of the free solvated cation is equal to that of ion-paired cesium thiocyanate 144 is consistent with all of the data. The other assumptions have only minor effects on the values of the ion—pair for— . O matlon parameters (Kip, AHip’ 61p 12 and 15—18 reveals that the parameters for the formation ). An inspection of Tables of ion—pairs are well determined and they are the same within their standard deviations, regardless of the model used. One obvious deficiency of the less critical assumption is that the chemical shifts obtained for the triple ions are very different from the values one expects on the basis of the overlap model. The origin of such behavior is not known to us. If the effect is real, and not merely a consequence of inadequacy of the model, the discrepancy must be caused by some kind of interaction between the two sets of overlapping orbitals in the triple ions, which will affect the average excitation energy in the Ramsey equa— tion (Equation 25). Another reason for such behavior may be due to changes in the short range ion—solvent interactions. When the anion or the cation interacts with the ion—pair, the solvent molecules may be rearranged in such a way that it causes a large chemical shift. In solvents such as methanol where ion-triple formation is not expected, the plots of 133Cs and 87Rh chemical shifts of 051 and RbI are not simple (Chapter V and Reference 211). The chemical shift changes rapidly at lower concentrations, but changes more slowly at high concentrations. 145 According to our model, at the lowest concentrations we studied,90% and 55% respectively of the cesium cations in cesium iodide and cesium tetraphenylborate are ion-paired at room temperature. At the highest salt concentrations (where the complexation by l8-crown-6 is usually studied) only 3% and 25% of the cesium cations in cesium iodide and cesium tetraphenylborate respectively are free. Therefore, in methylamine solutions,ion-association of cesium salts can- not be ignored in the study of the thermodynamicscfi‘the com- plexation of these salts by 18-crown-6. We have used the results of the ion-association of cesium salts in methyl- amine described in this chapter to obtain thermodynamic formation constants for complexation of cesium salts by 18-crown-6 in methylamine as discussed in Chapter IV. The conductance of cesium iodide in methylamine as a function of concentration was also measured in order to permit a comparison with the NMR results. This study will be des- cribed in Section 3 of this chapter. 3. Electrical Conductance Measurements of Cesium Iodide In Methylamine A. Introduction The thermodynamics of ion-association of cesium salts in methylamine has been investigated by 133Cs NMR tech- niques as described in the previous section. It was 146 pointed out that the thermodynamic parameters of the minor species cannot be well-determined from the NMR data. If we could independently determine the ion—pair formation constants of cesium salts in methylamine by another tech— nique, and use them as fixed values in treating the NMR data, we could obtain more information about further ion— association from our NMR data. Ion—pair association constants of cesium salts (Picrate, tetraphenylborate, and iodide) cannot be studied by UV spectroscopy, since the anion absorption band is either weak or insensitive to ion—pairing with the cesium cation (196). A common approach is to measure the electrical conductivity as a function of concentration, even though conductance theory faces difficulties in determining ion-association constants, as described in the Historical chapter. B. Results (i) Calibration of the Conductance Cell — The cell constant of 1.02l3i0.0004 cm-1 was determined by using dilute aqueous potassium chloride solutions and Equa— tion 1—32. Measured resistances of water and of potassium chloride solutions at 25.0°C are given in Table 22 at various frequencies. The following procedure was used for the calculation of the equivalent conductance. First, the resistance of the solvent at each frequency was computed from the parallel 147 moamo.fi mommo.fl Afilsovg O sso.o sam.o ss.o a me mowam ososm eoaxmomma.fi Awesov om maamm smmsm oofixmsmos.a oaomm oases comm omaam smmsm ooagaaoom.a weasw swsom . omom wawam mammm eoasmmoaa.a oammw oases mom mmwflm mzwmm woflxfimmmfi.fl mmsmm omsom sam mflwflm mmwmm moaxawmafl.a mmmmm omsom mam mamo.m mamm.fi Lopez popes + powwflmom AonV a soehsmom esseseom hostesses o m as OH x sz Hog eseesspm esm.oa exm.om AmEQOV oocopmflmom Spawmopom m50050< pct .ooao.oao.hm he hsoflosflom oeasoaso .soosz .soomamom eseespsm esm.om one so hoosssmahom .mm oases 148 resistance according to the equation l _ l. _ 1 (1 H, R ‘ "R 4‘ solv solv+SR SR where Rsolv’ RSR’ and Rsolv+SR are the resistances of the solvent, the standard resistor, and the solvent shunted by the standard resistor, respectively. Then the resistances of water and potassium chloride solutions were corrected Ifor the polarization effect at the electrodes and for the capacitance by-pass effect according to the equation, R = R + af2 + b meas 0 ;EE (1-33) The corrected resistances, R0, together with their marginal percent standard deviations are given in Table 22. The cell constant at each salt concentration was then computed from the corrected resistance according to the equation, ACRO _ x01 K _ R0 (3-9) 1000(1 - R KCl) 0 H2O Another set of calibrations was carried out at three dif- ferent concentrations of potassium chloride. The computed 149 cell constants were 1.02114, 1.02127, and 1.02125 cm'l. The average value of the cell constant from these five 1 determinations is 1.021310.0004 cm—. Cince the cell constant does not change over a concentration range of potassium 4 — 3.15 x 10‘” chloride (1.07 x 10- M), the procedure for the correction of resistances for the frequency dependence can also be used with confidence for the conductance measure- ments of cesium iodide in methylamine. (ii) Conductance of Cesium Iodide in Methylamine at -l5.7°C. - Resistances of cesium iodide solutions in methylamine were measured at -l5.7°C. The measured resistances at 617, 962, 2020, 3960 Ph: were corrected with the same procedure which was used for the calibration of the conductance cell. Equivalent conductances were then calculated from the following equation, = 1000K ( l _ l C Rosin Rosolv ) (3-10) and are given in Table 23. The plot of the equivalent con- ductance versus the square root of the concentration is shown in Figure 28. Since the specific conductance of methyl- 8 amine is very 1ow (1.1 x 10— ohm.1 cm—l), precise deter- mination of this quantity was not possible. The main problem in our experiment resulted from the use of a 150 Table 23. Equivalent Conductances of Cesium Iodide Solu— tions in Methylamine at -15.7°C. Conc. x 105 —1 A_l —l- . (M) (ohm cm eq ) 0A Weight 3.522 99.2 1.7 0.0015 9.022 70.7 1.4 0.0023 15.42 59.2 0.97 0.0047 22.87 52.3 0.65 0.011 31.02 47.3 0.38 0.031 49.28 39.23 0.18 0.13 73.34 33.48 0.10 , 0.42 103.3 29.137 0.058 1.25 142.0 25.715 0.041 2.5 200.5 22.189 ‘ 0.027 5.6 151 100 ' l D I 70 50 ~ - 40 - ~ sot - 20 . ~- A(.0."cm2 eq”) 1 I l 1.0 2.0 3.0' 4.0 5.0 . C’2 x 102(M’2) Figure 28. Plot of equivalent conductance versus the square root of the molar concentration of cesium iodide in methylamine at —l5.7°C. 152 glycol bath for the temperature maintenance. An oil bath would be preferable for such an experiment but was not available to us. An error analysis of the data was per- formed and the errors in the equivalent conductances were calculated and are also given in Table 23. The analysis shows that at low concentrations where the resistance of the solvent is not negligible compared to the resistance of the solution and also the relative error in the concentration is large, high standard deviations are associated with the calculated equivalent conductance. To take account of this error, appropriate standard deviations for equivalent conduc- tances and concentrations were used in the fitting of the data; i.e., each point was used with a weight proportional to the reciprocal of the variance for that point. The associated weights for each point according to the Onsager limiting law are also given in Table 23. C. Discussion Electrical conductance data of cesium iodide at -15.7°C were analyzed according to various conductance equations. The Onsager limiting law (LL) for weak electrolytes is expressed as, A = 01(AO - 3763) (1-21) and K = * The extended conductance equation which adds higher order terms to the conductance equation has the general linearized form (83), A=AO—S(Cd)l/2+EC log(Cd)+JlCd—KAA(Cd)yi—J2(Ca)3/2 (1—22) where S is the coefficient of the limiting law, and E de— pends only on the properties of the solvent and the charge on the ions, while J and J2 depend on the same parameters 1 and also on the distance of closest approach of ions. The coefficients E, J1 and J2 have different values according to the partiCular theory employed. Among the conductance equa— tions are: Pitt's equation (F) (77), the Fuoss—Hsia equa- tion (FH) (83) (both linearized by Fernandez—Prini (78,84)), the Fuoss-Hsia equation corrected by Chen (FHC) (197), and the Justice equation (J) (95,96). The Justice equa— tion consists essentially of setting the distance parameter equal to q (Bjerrum distance) in the FHC equation. The results of the analysis of data according to the above equa— tions are given in Table 24 and the values of the coef- ficients along with the subroutine EQN for the KINFIT pro— gram are given in Appendix 2. For all calculations a value of 5.3 A was selected as the distance parameter (Section 2B1) except for the Justice method in which the Bjerrum distance Of 33.0 A was used. Also the formation of the triple-ions was ignored since the mobilities of triple—ions are not known. _. _. .......__ . __. zoaxw.m Ama canoe Eogavmzz om.o mm.o omwammm moaxg:.mha.flv m 03.0 sm.fi H.mae.sm moaxflm:.ohos.mv ems 1.4 . my mm.o s.m ashamfi seasxa.aao.mv as mm.o o.H mmsmafi soasxo.flaa.mv a ese.o eso.o m.sho.aea soasxem.ohms.mv on <. Cm mHmSUHmom do A 1oo Eo 1E£ov A 12v compozom compofl>om mogosom do Ezm H mo H .M poopcmpm popnmfloz < m owopo>< .mcowpmsom cocopoSpcoo mSOHLm> an ooCfiopno Dow.mfl1 pm oQHEoncpoz CH opHUOH Esflmoo mo oocwpospcoo pcoflm>HSUm mCHpHEHq Ugo pcopmcoo compofloomm< map 00 monao> .qm magma 155 It is clear from 24, that both the ion-association constant and the limiting equivalent conductance depend strongly on the conductance equation used. The same behavior was observed by Gilkerson (198) for ion- association of cesium tetraphenylborate in acetonitrile, and of silver nitrate and lithium picrate in 2-butanone. As was mentioned in Chapter I, the apparent reason for such behavior is that both the higher order terms in the con— ductivity equation and the terms caused by association have the same concentration dependence in the first approximation. An increase in the sum ECdlog(C0) + J100 can be compensated for by a decrease in the ion-association constant. Our data clearly show that the most recent equations, which intro- duce higher order terms into the conductance equation, produce smaller ion-association constants as long as the distance parameter is the'same. On the other hand, changes in the distance parameter cause changes in the J1 and J2 coefficients. An increase in the distance parameter (Fuoss 1977 (199) and Justice (95)) causes J1 to decrease and J2 to increase. Both such changes in the J1 and J2 coefficients would be compensated for by an increase in the ion-association constant. D. Conclusion The ion-association constants obtained from conductance measurements depend on the conductance equation used for the 156 calculation of this parameter. The Onsager limiting law fits the data best as indicated by the smallest sum of the residual squares. The Justice method produces an ion- pair formation constant which is larger than that obtained from the NMR data. In this case, the fit is poor and the ion—pair formation constant and the infinite dilution equiva- lent conductance are not well-determined. Because the ion- association and the higher order terms in the conductance equation have the same concentration dependence, the separa- tion of these two effects does not appear to be possible with the available theory, at least at the level of precision of our measurements. Therefore the ion-association constants obtained from conductance measurements cannot be trusted until the separation of these two effects can be made or until an independent determination of A0 is made. 4. Comparison of NMR and Electrical Conductance Measurements The only value obtained for ion-pair association con- stant from the conductance measurement which is larger than the NMR value is that obtained by the Justice method. The ion—association constant obtained from other conduc- tance equations are smaller than the NMR values (Table 24). Electrical conductance measures the fraction of un- charged species. The equilibria in the solution are, K K + 1 2 M + x‘ z [M+.s.x‘] 2 M+.X- (3-11) 157 + _ where M and X are the solvated cation and anion respec- * + _ + _ tively, and M -S-X and M -X are solvent separated and contact ion-pairs. According to the above equilibria, the conductance ion—pair constant can be written as K = (MSX)+(MX)‘ <,_.2) cond (M)(X)Yi in which the charges are omitted for simplicity. Sub— stitution of the equilibrium constants into Equation 3-12 gives, Kcond = KlY§ It should be emphasized again that the limiting chemical shift at high concentration, 5MX’ is not the chemical shift of the contact ion-pair but is the population averaged chemical shift of the two types of ion-pairs. From the above discussion, it is clear that the ion- associations obtained from conductance and NMR measurement should be the same. The discrepancy between the calculated values (Table 2“) is probably due to the inadequacy of the conductance equations. OCHAPTER IV COMPLEXATION OF CESIUM SALTS BY l8-CROWN-6 IN METHYLAMINE AND LIQUID AMMONIA 159 I. Introduction The complexation of the cesium cation by lB-crown-o (18C6) in aqueous and methanolic solutions as well as in mixtures of these solvents has been studied by potentiometric (160) and calorimetric (173,17A) techniques. Mei et al. (200, 201) used 13305 NMR to study complexation of cesium tetra- phenylborate by 18—crown—6 in six nonaqueous solvents. In all solvents studied, the formation of the 1:1 complex was followed by the addition of a second molecule of the ligand to form a "sandwich" 2:1 complex. The use of high donor solvents such as liquid ammonia and methylamine to study alkali metal solutions in the pres- ence of macrocyclic ligands (183) motivated us to investi- gate the complexation of cesium salts by lS-crown-6 in these solvents. In the course of such studies with methyl- amine, it became clear that many equilibria are involved in the complexation process. Therefore an extensive study of the system was required to obtain the thermodynamic. formation constants of the 1:1 and 2:1 complexes. The results of these studies are presented in this chapter and compared with the results in liquid ammonia and in other solvents. 160 161 2. Complexation of Cesium Salts by 18-crown-6 in Methyl- amine A. Mole Ratio Dependence of l33Cs‘Chemical Shift in Methylamine (1) Results - Cesium-133 chemical shifts of cesium iodide and cesium tetraphenylborate in the presence of 18-crown-6 were measured as a function of (18-crown—6)/ (Cs+) mole ratio (R) at a fixed concentration of the salt in methylamine solutions. The results for cesium iodide at various temperatures are given in Table 25 and shown in Figure 29. At all temperatures, an upfield shift resu1ts as the mole ratio increases. This means that the cesium cation interacts more strongly with the iodide ion than with lB-crown-6. A plot of the chemical shift versus mole ratio is practically linear in the range of O < R < l. The linearity of the plots below R = l indi- cates the formation of a strong lzl complex (Kf Z 10“). The formation of the strong l:l complex is followed by the formation of a weaker 2:1 complex as indicated by a change in the slope of the mole ratio plot above R = 1. Similar results for cesium tetraphenylborate are given in Table 26 and illustrated in Figure 30. A 162 Table 25. Mole Ratio Study of 18C6, CsI Complexes in Methyl— amine at Various Temperatures; (Cs+) = 0.0206 1 0.0008 M- 60bS(ppm) (¥8ég)§?0:$) +25.1°C +13.2°C +9.5°C +6.0°C 0.0 127.10 127.10 126.36 126.13 0.205 117.91 118.22 117.52 117.37 0.365 109.85 110.08 109.69 109.30 0.070 100.30 . 100.73 100.26 100.26 0.576 100.85 101.16 100.62 100.50 0.810 89.05 89.10 88.52 88.37 0.885 87.13 86.58 86.35 85.88 0.958 80.02 83.87 83.02 82.63 1.012 82.70 80.92 80.38 79.99 1.197 79.99 78.13 . 77.u3 76.73 1.569 75.57 72.63. 71.15 69.83 2.050 70.80 66.50 60.25 62.39 2.295 69.00 60.60 61.39 59.91 2.668 65.80 59.29 56.81 50.17 3.170 63.32 50.95 51.92 08.82 0.531 50.30 03.55 39.59 35.87 6.096 07.58 35.09 30.29 26.25 .11.109 28.57 13.27 8.30 0.23 73.66* 17.72 2.37 ____________ 163 Table 25. Continued. obs Figure 29. Cesium-133 chemical shift versus (l8-crown-6)/ (CsI) mole ratio and temperature in methylamine; (CsI) = 0.02 g. 166 Table 26. Mole Ratio Study of 1806, CsBPhu Complexes in Methylamine at Various Temperatures; (Cs+) = 0.0108 0.0005 g. 60bS(ppm) Mole Ratio Temperature, °C (1806)/(Cs+) 25.0 13.2 5.8 —3.0 0.000a -7.05 -3.35 —1.65 0.25 0.209 -3.06 0.90 2.60 0.50 0.095 3.07 6.97 8.73 10.82 0.600 6.01 9.03 11.52 13.50 0.857 9.89 13.50 15.71 17.80 0.875 11.29 10.78 16.33 18.30 0.901 11.68 10.78 16.08 18.50 0.922 11.68 10.78 16.79 18.89 0.999 12.05 15.78 17.57 19.08 1.089 11.68 10.39 16.02 17.72 1.365 10.28 11.21 11-00 11.36 1.511 8.57 8.81 8.02 7.11 1.516 7.33 6.90 6.25 0.39 1.639 6.63 6.03 0.85 2.21 2.013 3.30 1.00 -0.73 —0 30 2.377 -0.65 -0.10 -6.90 -ll.59 2.720 -2.98 -7.00 -10.35 -15.63 3.100 —6.07 -11.51 —15.23 -20.66 3.331 —8.02 -13.29 -17.02 —22.99 3.797 -10.97 -17.09 —20.66 -26.55 6.005 —19.96 -26.00 -29.81 -30.85 11.076 -28.96 —33.92 -36.79 —00.51 167 Table 26. Continued. 60bS(ppm) Mole Ratio Temperature, °C (18C6)/(Cs+) —16.2 —32.1 _ -07.7 0.000a 1.65 0.209 6.25 b b 0.095 12.91 b b 0.600 15.86 b b 0.857 20.36 23.77 b 0.875 20.82 23.85 b 0.901 21.06 20.39 b 0.922 21.60 25.17 b 0.999 22.57 25.36 b 1.089 19.82 22.38 b 1.365 9.97 8.65 7.95 1.511 3.80 0.28 —2.50 1.516 0.51 -3.68 -7.17 1.639 -2.83 -8.56 -10.57 2.013 -1l.28 -19.19 -27.60 2.377 -20.58 -29.19 -38.19 2.720 -25.00 -33.77 -0O.9O 3.100 -31.13 -38.11 —00.02 3.331 -32.68 -39.82 -05.08 3.797 -35.90 -01.83 —06.60 6.005 -0l.75 —05.79 -08.50 11.076 -05.50 -07.09 -09.16 aObtained from extrapolation. b Precipitation in solution. 1 1 l 1 1 1 1 1 1 | 2 3 44 5 6 7’ 8 9 K) H Mole Ratio (I8-crown-6)/(CsBph4) _——, Figure 30. Cesium-133 chemical shift versus (18-crown—6)/ (CsBPhu) mole ratio and temperature in methyl— amine; (CsBPhu) = 0.01 M. 169 linear downfield shift followed by an upfield shift which gradually approaches a limiting value can be explained by the formation of a relatively strong 1:1 complex followed by the addition of a second molecule of the ligand to form a 2:1 sandwich complex. 133 (ii) Discussion — The variation of the Cs chemical shift as a function of (18-crown—6)/(Cs+) mole ratio in methylamine can be explained by the formation of a strong 1:1 complex followed by the formation of a 2:1 complex. If we were able to neglect ion—pair formation for the salt, the 1:1 complex, and the 2:1 complex, the data could be analyzed according to the equilibria, M+c (0—1) M+C <0—2> in which M+, C, M+C, M+C2 are the cesium cation, the ligand, the 1:1 complex, and the 2:1 complex, respectively. Since the mole ratio plot is linear below mole ratio of unity, the formation constant of the 1:1 complex cannot be ob- tained from the data and only a lower limit of this value can be estimated (Kl : 100). Then the variation of the chemical shift above R = 1 could be used to obtain the formation constant and the limiting chemical shift of the 170 2:1 complex by assuming that the formation of the 1:1 com- plex is complete at a mole ratio of unity. The solution to these equations and the subroutine EQN for use with KINFIT Euwagivenelsewhere (186). The results,according to this simple scheme,are given in Table 27 for cesium iodide and in Table 28 for cesium tetraphenylborate. The enthalpies and entropies of complex formation were obtained by using the KINFIT program. The van't Hoff plots are shown in Figures 31 and 32. In the 2:1 sandwich complex, the cesium cation is ex- pected to be effectively isolated from the solvent and the counter-ion. However, at least for the iodide salt, the limiting shift is strongly temperature dependent. In addition,comparison of Tables 27 and 28 shows that the formation constant of the 2:1 complex at a given temperature is strongly anion dependent. Both of these effects suggest that other equilibria are important in the solution. We have seen in Chapter III that cesium salts are highly associated in methylamine. Therefore it is reasonable to assume that the 1:1 complex,in which the cesium cation sits above the 18-crown-6 cavity,can also form ion-pairs with the anion. Since the ion-association constant of cesium iodide is stronger than that of cesium tetraphenylborate (Chapter III), the calculated complexation formation constant accord- ing to equilibrium (0-2) should be smaller for cesium iodide than for cesium tetraphenylborate, because the 171 Table 27. Thermodynamic Parameters for the Formation of the 2:1 Complex of l8—Crown—6 and CsI in Methyl— amine. K i 100. t°:O.5 1:21 (611m)2:l (0C) - (M ) (ppm) 25.1 5.0010.20 —25.8:2.0 13.2 6.93:0.13 -36.8:1.0 9.5 7.87i0.19 -39.9i1.3 6.0 8.82i0.17 d01.2i1.0 .2 10.30i0.15 -00.5i0.7 -2.5 12.68i0.29 —01.3iO.9 -10.3 17.29:0.00 —00.2i0.8 -16.3 22.71:0.95 -05.5i1.2 -32.1 00.12i0.98 —50.2iO.5 -00.0 65.3:2.2 -09.3i0.0 —08.0 75.5 11.9 —50.0:0.3 AG398°= —0.9610.02 kcal.mo1e-l. AHO = —5.210.1 kca1.mole_l. A80 = -l0.2i0.0 e.u. 172 Table 28. Thermodynamic Parameters for the Formation of the 2:1 Complex of l8—Crown-6 and CsBPhu in Methylamine. K1 1 10 . t°i0.5 "¥_" 1 (611m)2:1 (00) (M ) (ppm) 25.0 23.1:1.0 -07. i .9 13.2 03.6:1.5 -05.8iO.5 5.8 58.9:1.0 -06.0:0.3 -3.0 89.7il.9 -07.6iO.2 -16.2 186.9i0.1 -09.3i0.1 -32.1 007 i12 -09.5i0.1 -07.7 893 i09 -50.1i0.1 6039800=-1.86:0.03 kca1.m61e‘l. AH° = -6.8310.25 kca1.mole . 880 = -16.35i0.96 e.u. 173 4.2 P 3.8 - 3.4 *- 3.0- an2 2.6 " 2.2 " l.8- lg; l I l I l l ' 3.2 3.4 3.6 3.8 4.0 4.2 4.4 l/T x lo3 (°K"') Figure 31. Ln"K2" vs 1/T for the 2:1 complex of 18-crown- 6 and CsI in methylamine. K1 1 10“, (Cs+)=0.02M. 170 20 £55" 5£P' In K2 ‘45- 135- I I I I I L 303.2 34 3.6 3.8 40 4.2 44 l/T x 10’ PK") Figure 32. Ln "Ké'zs l/T for the 2:1 complex of l8-crown-6 + and CsBPhu in methylamine. K1 1 10“, (Cs )=0.01M. “-5. ligand and the anion are competing for the cesium cation. Thus, although the simple picture given above can be used to fit the data,it is obvious that ion—association of the 1:1 complex cannot be ignored in the treatment of data. It will be shown later in this chapter that even the 2:1 complex forms ion—pairs in methylamine (which are probably not contact ion—pairs). The complete scheme for this system can be written as, K _ ta _ _ K + x z x .M+.X + — ip — M + x 2 M x K t _ + M+ : M+ x .M+ + + c 0 6C ++ ++ KX K _ A _ MC + x : MC+.X (0-3) + + c c KC2 ++ ++KX2 K MG; + x' gf2 MG;.X_ in which MC+, M05, Mc+.x‘, and MCZ.X‘ are the 1:1 complex, 2:1 complex, ion—paired 1:1 complex and ion—paired 2:1 complex, respectively. Other symbols have the same mean— ings as in Chapter III. It is important to note that the 1:1 complex might form both contact and solvent—separated 176 ion-pairs but the separation of the formation constants of these two kinds of ion-pairs is not possible from the NMR data (see Chapter III). It should also be obvious from the complexity of the complete scheme that the evaluation of all the equilibrium constants as functions of temperature and the chemical shifts of all the species is an impossible task. Clearly it will be necessary to make some assumptions and approximations in evaluating the parameters. The concentration dependence of the chemical shift of the 1:1 complex could, in principle, be used to evaluate K c’ KA’ and Ki as functions of temperature. Then the com- plexation formation constant and the ion-pair formation constant of the 2:1 complex might be obtained from the mole ratio studies by using proper equilibria. The results of studies of this type are given in the next sections. 133 B. Concentration Dependence of the Cs Chemical Shift of the 1:1 Complex in Methylamine (1) Results — The concentration dependence of the 133CS chemical shift of the 1:1 complexes of 18-crown-6 with cesium salts was examined. The results at various temperatures for cesium iodide, cesium tetraphenylborate, and cesium thiocyanate are given in Tables 29, 30, and 31, respectively. Plots of the 133Cs chemical shift versus the concentration of the 1:1 complex are shown in Figures 33, 30 and 35 for each salt. Table 29. Concentration Dependence of the l‘33Cs Chemical Shift of the 1:1 Complex of CsI and 1806 in Methylamine at Various Temperatures. 6obs (ppm) 0 Cone. Temperature C (M) 25.2 12.3 6.0 —2.1 —10.0 -15.9 0.00005 95.81 ————— 79.53 76.35 70.72 0.00092 91.00 {85.50 83.95 80.85 79.37 78.29 90.85 85.73 78.21 91.16 0.00202 89.30 85.11 83.97 82.09 80.77 79.68 79.99 0.00377 87.98 80.60 83.08 82.09 81.39 80.69 87.98 0.00070 87.36 ————— 83.08 81.93 81.39 80.85 0.00606 86.66 80.02 83.00 82.20 81.78 81.15 86.76 83.09 0.00755 86.03 .83'95 83.09 82.32 81.93 81.07 0.00960 85.73 83.79 83.17 82.32 81.93 81.62 0.01198 85.02 83.00 82.86 82.16 81.93 81.62 85.19 0.01793 80.57 83.02 82.71 82.20 82.09 81.72 178 Table 30. Concentration Dependence of the 133Cs Chemical Shift of the 1:1 Complex of CsBPhu and 18C6 in Methylamine at Various Temperatures. 6 obs Conc. Temperature (M) 25.0 6.0 -2.5 —16.1 0.00020 20.90a ---------- 30.36a 0.00051 17.18a 23.38a 25.00a 28.25a 0.00100 15.63 21.60 23.62a 26.80a 0.00158 10.50 20.00 22.38 26.02a 0.00206 10.00 19.82 21.60 25.55 0.00272 13.85 19.51 21.75 20.07 0.00002 13.30 18.73 21.06 23.85 0.00090 12.91 18.30 20.20 23.62 0.00750 12.60 18.03 19.70 23.07 12.30 0.01129 {12.05 17.65 19.12 23.07 aThese points were omitted in the analysis of the data (see text for explanation). 179 Table 31. Concentration Dependence of the 133Cs Chemical Shift of the 1:1 Complex of CsSCN and 18C6 in Methylamine at Various Temperatures. dobs (ppm) Conc. (M) Temperature 25.0 6.0 -16.0 -30.8 -50.0 0.00099 03.23 01.10 00.21 00.37 01.99 0.00198 {00.52 39.28 39.36 00.29 02.23 00.21 0.00503 37.88 37.73 {38.82 00.37 02.31 39.13 0.00760 36.09 37.30 38.50 00.00 02.55 0.00993 36.02 36.80 38.58 00.52 02.38 0.01273 35.71 36.01 38.55 ----- 02.69 0.02010 35.09 36.01 38.01 00.68 02.15 180 72....... 76 80 84 80bs (ppm) 88 92 96‘ 1 I I I l000.0 0.4 0.8 l2 L6 2.0 ' [65*] x lo2 (M) Figure 33. Concentration dependence of the 133Cs chemical shift of the 1:1 complex of 18—crown—6 and CsI in methylamine at various temperatures. 181 l2- 25.0°C '6.— A ' ‘ aco" g o “25°C 3 20- U) 7. CI‘O .0 o -l6l°c 0° 243° , 0 V " V 28" 32 I J 0.0 0.4 0.8 l.2 [Cs’] x IO2 (M) Figure 30. Concentration dependence of the 133Cs chemical shift of the 1:1 complex of 18—crown—6 and CsBPhu in methylamine at various temperatures. 182 34 j 7 l r I l 250°C 60°C ’7 462°C 6 6C? 5'308°C v ‘504°C 44 ‘ 4 ' l g 1 0.0 0.4 0.8 l2 l6 2.0 2.2 [65*] x lo2 (M) Figure 35. Concentration dependence of the 133Cs chemical shift of the 1:1 complex of 18—crown—6 and CsSCN in methylamine at various temperatures. 183 (ii) Discussion — (a) General Discussion - In all cases the variation of the chemical shift as a function of concentration and temperature reflects the competition between ion-pair formation and complex formation. This behavior is most pronounced in the cesium iodide case. The cesium iodide data (Figure 33) show that at low concentra- tions and high temperatures the chemical shift approaches that of the Cs+.I- ion-pair (120 ppm), while at low concen- trations and low temperatures an upfield shift occurs as concentration decreases. It seems reasonable to believe that this upfield shift is due to the formation of the MC+ complex. At high concentrations all curves converge to the same chemical shift whichijsbelieved to be the chemical shift of the ion-paired complex (MC+.I-). The concentration and temperature dependence of the chemical shifts of the 1:1 complexes of 18—crown-6 with cesium tetraphenylborate (Figure 30) and cesium thiocyanate (Figure 35) can also be explained in the same way. However, the cesium iodide data are more illustrative since the chemical shift of the ion-paired complex, 6 lies between the chemical shift of the MC+.X" ion—paired salt, 6 and the Chemical Shift of the M+.x-’ complex, 6MC+° According to the above discussion, the equilibria responsible for the chemical shift changes of the 1:1 complexes as a function of concentration and temperature can be written as, + _ M .x M+ 7 + Kt M+ + c c - (0-0) The formation constants and the chemical shifts of the ion- pairs and triple ions as well as the chemical shift of the free cation are known (Chapter III) and can be used as con- stants in the above equilibria in order to obtain other param- eters. The solution to the above equilibria is given in Ap- pendix 3A together with the subroutine EQN for use with KINFIT. To completely analyze the data for each salt at least six parameters would have to be adjusted. These are 0 and 6 + O O o . KX, AHX, KA’ AHA, MC+’ MC .X” Since KC is a depen— dent variable (KC = K Kx/KA) it need not be independently 1p adjusted. Again, as in the case of cesium salts in methyl- amine (Chapter III), the determination of all parameters is not possible since some of the parameters are highly correlated with each other and/or correspond to minor species. For example, the chemical shift of the com— plex, 6MC+’ is difficult to determine from the data. Since ion-pair formation of both the salt and the 1:1 complex are competing with simple complex formation, so that the concentration of MC+ in solution is small at all con— centrations and temperatures. Although this species has its highest mole fraction at the lowest concentration and the lowest temperature, even under these favorable circum— stances, it amounts to less than 15% of the total cesium concentration. The ion—pair formation constant of the 1:1 complex, KA, and its enthalpy of formation, AHE are strongly coupled and the simultaneous determination of these param— eters is not possible. Consequently, we were forced to fix 6MC+ and AHE (or KA) at "reasonable" values in order to obtain the other parameters. The rationale behind our choice Of 6MC+ and AHK will be described in the next two sections. (6) The chemical shift of the 1:1 complex, 6MC+° As described in Chapter III, the chemical shift of cesium thiocyanate in methylamine is independent of concentration so that we were unable to determine the ion—association parameters for this salt. Therefore, it might seem that the concentration dependence of the chemical shift of the 1:1 complex of 18—crown-6 with CsSCN would not provide 186 much information about the complexation reaction. On the contrary, however, the most valuable information about the chemical shift of the 1:1 complex, can be obtained directly from the cesium thiocyanate data (Figure 35). Since the chemical shift of cesium thiocyanate is in- dependent of concentration and is presumably the same as the chemical shift of the free cesium cation (see Chapter III), the chemical shifts of the 1:1 complex, 6MC+’ and of the ion-paired complex, 5 should also be the Mc+.x-’ same. A closer look at Figure 35 shows that only the chemical shift of the 1:1 complex (and the ion-paired com— plex) can be responsible for the upfield shift at high temperatures, because the chemical shiftscfl‘other species, igeg, the free cation and the ion-pair, are both larger than all of the observed chemical shifts. Therefore, extrapolation of the curves to high concentration at each temperature should provide the chemical shift of the 1:1 complex (and the ion-paired complex). Fortunately, the curves at low temperatures level off at high concentrations and 5MC+ can be obtained directly. If we assume that the chemical shift of MC+ is linearly dependent on temperature, (as is the case for ion—pairs and the free ions), then 6 + at high temperatures can be determined from the ex- MC trapolation of the low temperature data. In fact a plot of the chemical shift versus temperature is almost linear at low temperatures (Figure 36). By complete analysis of 187 Figure 36. Limiting 133Cs chemical shift of the 1:1 com- plex of 1806 and CsSCN vs temperature in methyl- amine. A experimental values at the highest concentration studied (Figure 35), 0 calculated values obtained from the parameters of Table 36. 188 the data (section iiic) we found that the chemical shift of the complex at a given temperature can be expressed as, (61%.)t = 32.5 —o.13 (t - 25°C) (0—5> where the chemical shift of the complex at 25°C is taken as the reference chemical shift. Now let us see if the above conclusion about the chemical shift of the 1:1 complex is consistent with the other results. The cesium iodide data (Figure 33) suggest that the chemi- cal shift of the complex at 25°C should be smaller than the limiting chemical shift, 6MC+ I‘ z 80 ppm, since the cesium cation interacts more strongly with the iodide ion than with the solvent. The same data give an upper limit of 72 ppm for 6MC+ at —l6°C. The cesium tetraphenylborate data (Figure 30) indicate that the chemical shift of the complex is greater than 12 ppm at 25°C and greater than 23 ppm at —16°C, because the cesium cation interacts more strongly with the solvent than with the tetraphenylborate anion. Therefore, the chemical shift of the complex is limited to, 80 > 6 > 12 ppm at 25°C Mc+ _ O 72 > 6MC+ > 23 ppm at 16 C which is consistent with the above conclusion. 189 (c) Enthalpy of formation of the ion-paired complex As stated earlier, the ion-pair formation constant of the 1:1 complex, KA’ and the enthalpy of the formation of the ion- paired complex, AH: simultaneous determination of these two parameters is not , are highly correlated and therefore possible. This is evident from attempts to obtain these two parameters simultaneously from the cesium iodide data. The results are given in Table 32. The standard deviation associated with AHO A is even greater than the parameter itself. is very large and the standard devia- tion of KA This indicates that the information content of the data is insufficient to determine both parameters and we must fix one of those at a reasonable value. + - The free energies of formation of the ion—pair, M .X , + _ and the ion—paired complex, MC .X can be written as, AGip RTanip AHip IASip (0 6) and 0 = - = 0 _ 0 _ AGA RTanA AHA TASA (0 7) If we assume that the difference in the free energies of ion-association of the free cesium cation and the complexed cation is entirely of enthalpic nature, then we can write, 0 _ o _ o = _ 6AS — ASA ASip 0 (0 8) 190 Table 32. Thermodynamic Parameters forthe Formation of the 1:1 Complex of l8-Crown-6 and CsI in Methylamine at 25.0°C; with Adjustment of AHO. A xx (8.53il.65)x103 ' M'l AH; —20.0:3.1 kcal.mole-l KA (0.3:3.6)x105 M-l 6H3 7.1:3.1 kca1.mole-l 6MC+.X- 83.08:0.20 ppm -1 bMC+ X‘ -0.002:0.011 _ ppm.deg 56(3) 0.30 ppm aAverage standard deviation of chemical shifts (the symbol has the same meaning in subsequent tables). 191 Subtracting Equations 0—6 and 0—7 and introducing 5AS° from Equation 0—8 gives, K. 0= 0 1p AH AHip + RTln (KA ) (0-9) With this assumption it is only necessary to adjust one of the parameters (KA or ARK) since Kip and Ang are known from the salt results (Chapter III). iii. Analysis of the Data The concentration dependences of the chemical shifts of the 1:1 complexes of l8—crown—6 with cesium salts were analyzed according to equilibria (0-0). Since the data for each salt do not provide equal amounts of information, the treatment of the data for each salt will be discussed separately. The solution to the equilibria and the general subroutine EQN are given in Appendix 3A. (a) Cesium iodide — The concentration and tempera- ture dependence of the chemical shift of Cs+l8C6.I_ was analyzed according to equilibria (0-0). The ion—association parameters of the salt were taken from Table 12. Since the values of these parameters are internally consistent, the choice of the model for ion-association should not have an appreciable effect on the complexation parameters. The chemical shift of the complex at various temperatures was 192 computed from Equation 0-5. The enthalpy of formation of the complex was obtained from Equation 0.9. In addi- tion it was considered that the chemical shift of the ion- paired complex, 6 is linearly dependent on tempera- MC.X’ ture according to, (6 ) = (6 MC.X t + b(t - 25°) (0-10) MC.X)25° The calculated parameters are given in Table 33. It should be noted particularly that the ion—pair formation constant of the 1:1 complex is similar to that of the uncomplexed salt. If the cesium cation were forming primarily contact ion-pairs in the salt solution, then one would expect that the ion-pairs would dissociate upon the addition of the ligand and the ion-pair formation constant of the complexed cation would be significantly smaller than that of the salt. Therefore it is reasonable to conclude that even the uncomplexed salts form largely solvent separated ion-pairs. Shchori et al. (167) also found that the dis- sociation constants of NaBPhu (Kd = 5.02 x 10'5 M at 20°C) and Na+DBl8-crown-6.BPhu- (Kd = 6.00 x 10'5 M at 20°C) in dimethoxyethane are of the same order of magnitude, which is an indication of "loose" ion—pair formation. In addition Boileau et al. (172) investigated the ion— association of Na+BPhu_ and K+BPhu- and their complexes with cryptands in tetrahydrofuran conductometrically. Table 33. 193 Thermodynamic Parameters for the Formation of the 1:1 Complex of 1806 and CsI in Methylamine at 25.0°C; with Calculation of AHX from Other Adjustable Parameters. Fixed Parameters = _ _ O (6MC)t 32.5 0.13 (t 25 ) O = O T AHA AHip + RTln(Kip/KA) Adjustable Parameters KA (1.51:0.06)x105 xx (6.33:0.00)x103 AH§ -16.00:0.53 SMC.X , 82.69i0.2l bMC.X -0.006:0.007 5 0.30 5 K K 0 Kc 1; X (1.07:0. l5)x10 A 0 - TO 0 AHC — Ahip + AHX .. AHO A -16.72:0.80 (From Equation 0—5) (From Equation 0-9) kca1.mole-l ppm ppm.deg-l ppm M-1 kcal.mole-l The authors found that the dissociation constants of Na+— BPhu‘ (Kd = 9.33 x 10"5 M at 20°C) and Na+22l.BPh4_ (Kd = 8.97 x 10_5 M at 20°C) are almost equal. Similarly the dis— sociation constants of K+.BPhu~ (Kd = 0.39 x 10_5 M at 20°C) and K+222.BPh4_ (Kd = 8.16 x 10‘5 M at 20°C) are of the same order of magnitude. The authors concluded that the un— complexed salts form solvent separated ion pairs in tetra— hydrofuran. In summary, our results are consistent with a model in which both salt and complex form largely "loose" ion—pairs in methylamine and are in accord with the results of two sets of authors for similar systems. However, the effect of the anion on the chemical shift shows that some contact ion— pairs must be formed. (b) Cesium tetraphenylborate — The concentration and temperature dependence of the 1:1 complex of l8—crown—6 and cesium tetraphenylborate in methylamine was analyzed ac— cording to equilibria (0—0). The ion—association param— eters from Table 12 were used as constants. The chemical shift of MC+ was obtained from the limiting chemical shift of the CsSCN data as described above (Equation 0—5) and a linear temperature dependence was assumed for the chemical shift of the ion-paired complex (Equation 0—10). The value of AH: was fixed at the value of Ang. The results obtained in this way are given in Table 30. In the analysis of the Table 30. Thermodynamic Parameters for the Formation of the 1:1 Complex of 1806 and CsBPhu in Methyl- - o.- 0= 0 amine at 25.0 C, With AHA AHip. Fixed Parameters (6M0)t = 32.5-0.13 (t—25°) (FrOm Equation 0—5 ) AHX = Ang Adjustable Parameters KA (l.30:0.20)xlou M“l KX (1.08:0.63)x100 M‘l AH; —8:25 kca1.mole_l 6M0.X 8.08:1.10 ppm bMC X —0.l2i0.02 ppm.deg'l 66 0.22 ppm , _..—_—__L 196 data we had to omit some of the points at low concentra— tion, since it was not possible to achieve convergence on the concentration of species. The solution to the equilibrium equations were compli- cated (Appendix 3A), therefore, the iteration procedure for the calculation of the concentrations was based on the major equilibria in the solution and successive corrections for the concentration of the minor species, M+, X_.M+.X—, and M+.X—.M+. At low concentrations and especially at low temperatures the fraction of M+ is not small (3 10%) and therefore the above procedure would not work for these points. However, neglecting these points should not have signifi— cant effect on the calculated parameters since it is only a result of the convergence failure when these concentrations are included and does not depend on the adjustable parameters. The only effect of neglecting these points is on the cal— culated standard deviations of the parameters. Table 30 shows that the standard deviation in AH; is huge. The reason is that this parameter is largely determined by points at low concentrations and temperatures, some of which had to be discarded. To improve the standard devia— tion in AH; we can use the parameters obtained for CsI. The formation constant of the unassociated complex, Kc’ is independent of the counter ion, + K M +0 :PMC+ 197 The value of KC can be calculated from cesium iodide data (KC = Kipr/KA) and then used as a constant in the tetra- phenylborate data. The enthalpy of the formation of the unassociated complex can also be expressed in terms of the enthalpies of formation of other species (AH: = AH; + p AH; - AHX). Thus only four parameters must be adjusted. O 0 These are Kx’ AHX, 6MC.X’ and bMC.X' Then KA and AHA can be expressed in terms of the other parameters. The calculated parameters are given in Table 35. These results are more meaningful than the results in Table 30 since KC (which is also strongly dependent on the low concentration points) has been used as a constant. The use of K0 and AH: from the cesium iodide data to fit the cesium tetra— phenylborate data has another advantage in that it provides a check on the consistency of the two data sets. The ion— pair formation constant of the complex for cesium tetrae phenylborate is the same order of magnitude as the ion- pair formation constant of the salt and the enthalpies of formation of these Species are almost equal. This sug- gests again that "loose" ion-pairs are important as was the case with cesium iodide. The value of KX is somewhat smaller for CsI than for CsBPhu which reflects the different degrees of ion-pair formation of the two salts. (c) Cesium thiocyanate - Since the chemical shift of the free cesium cation and ion-paired cesium thiocyanate 198 Table 35. Thermodynamic Parameters for the Formation of the 1:1 Complex of 1806 and CsBPhu in Methyl- amine at 25.0°C; K0 was used as a Constant. Fixed Parameters (6M0)t = 32 5—0.l3 (t—25°) From Equation 0—10 KC = 1.07x10“ M‘l From Table 30 AH: = —16 72 Kcal.mole‘l From Table 30 Adjustable Parameters KX (8.00:1.0)x103 M"l AH; —l8.80i0.95 kcalmole’l —l bMC.X —0.l9:0.03 ppm.deg 66 g 0.23 ppm K. K KA = —iE—§ (l.l6iO.35)XlOu M‘1 K C o __ o o_ o ‘1 AHA — AHip + AHX AHC 2.63:1.31 kca1.mole 199 are the same, the ion—pair formation constant of this salt cannot be obtained from the concentration dependence of the chemical shift. Similarly the ion—pair formation constant of the complex, KA’ cannot be extracted from the concentra— tion dependence of the chemical shift of the 1:1 complex, be— cause the chemical shifts of the complex and ion—paired complex are the same. We have shown in Chapter III that in methylamine solu— tions CsSCN is ion-paired. The value of KX can be approxi— mated by considering that cesium thiocyanate is completely ion—paired in methylamine. Then the only equilibrium which causes changes in the chemical shift would be, K M+.X‘ + c 33‘ MC+.x‘ Four parameters were adjusted to analyze the data. These are KanHE’ 6MC+.X" and bMC+.X"' The results are given in Table 36. The other extreme case is that cesium thiocyanate is com- pletely dissociated in methylamine although this would be extremely unlikely being given the low dielectric constant Of the medium. In such case, the only equilibrium respon- sible for the chemical shift would be, The parameters obtained for this extreme case are the same as the parameters of Table 36 except that KC==KX and AHg==AH;. 200 Table 36. Thermodynamic Parameters for the Formation of the 1:1 Complex of 18C6 and CsSCN in Methylamine at 25.0°C. Assumption a) CsSCN is completely ion-paired in methylamine then the only equilibrium involved is K - X _ M+.X + c z MC+.X b) CsSCN is completely dissociated in methylamine K C M++C2MC+ Adjustable Parameters Kx or KC = (0.87:0.53)x103 M‘l AH; or AH% = —13.50t0.73 Kcal.mo1e‘l — i 5MC.X or 6MC - 32.51 0.23 ppm b or b = -0.130i0.000 ppm.cieg‘l MC.X MC The fact that Kg obtained by assuming complete dis— sociation of the salt is different from the values ob— tained from data for the other salts also confirms that cesium thiocyanate is associated in methylamine. An at— tempt was also made to calculate the parameters of Table. 36 by fixing KA and Kip at various values. The values for K and AH; calculated in this way did not change appre— ciably. Since KX for CsSCN is smaller than the correspond— ing values for the other salts, it can be concluded that CsSCN is more associated than either cesium tetraphenyl— borate or cesium iodide. The ratio of the ion—pair forma— + _ + _ tion constants of Cs .SCN and Cs 18—crown-6.SCN can be ap— proximated and has a value of 2.2 (Kip/KA = Kc/Kx z 1.07 U x 10'/0.87 x 103). This ratio for cesium iodide and cesium tetraphenylborate is 1.67 and 1.27 respectively. (d) Summary — The thermodynamic parameters for the formation of the 1:1 complex between 18—crown—6 and cesium salts in methylamine at 25.0°C are summarized in Table 37. c AHg It is interesting to note that MC+) is enthalpy stabilized 747251 The complex formation (M+ + C and entropy destabilized. with very few exceptions, macrocyclic complexes in non— aqueous solvents seem to be enthalpy stabilized but entropy destabilized. It would be very helpful to be able to describe the solvent dependence of AHg and ASE of complexation, but the 202 Table 37. Thermodynamic Parameters of the Complexation of Cesium Salts by l8—Crown—6 in Methylamine at 25.0°C. . Salt (a) Parameter CsI CsBPhu CsSCN Kip (M'l)' 2.50 x 105 1.01xlOu >105 AH§p(Kcal.mole‘l) 3.86 0.7 ASS:p (e.u.) 37.7 30.8 KA (M‘l) 1.51 x 105 1.16x10” >105 AHZthis model which is more than the experimental error. Since equilibrium (0-12) does not contribute to the chemical shift changes, approximate values for Kx’ AH; and 6MC X could be obtained from the 2 analysis of the data according to equilibrium (0-11) alone. Table 38. Thermodynamic Parameters for the Formation of the 2:1 Complex of CsI with 18—Crown-6 in Methylamine at 25.0°C; Assuming Complete Ion- Pair Formation of the MC+ Complex at (Cs+)=0.02M. Assumptions: 1:1 complex is completely formed and completely ion— paired at 0.02 M total cesium iodide concentration. K. AHO = AH° + RTtn<~lB) A2 ip KA2 6 = 6 MC2 MC2X Adjustable Parameters (from Equation 0—11) —1 KX2 = 0.52iO.20 M AH§2 = —5.03iO.l9 kca1.mole—l KA2 = (0.6:13) x 105 M"1 GMCZX = —06.8il.07 ppm 65 = 2.0 ppm 1__131 208 The results for CST are given in Table 39a. The average standard deviation in the chemical shift is also 2.0 ppm according to this analysis. In order to examine whether this high standard deviation is due to the neglect of the MC+ concentration, we improved our model by assigning an average weighted chemical shift, 8 to the chemical shift MC+.x" + - of MC .X at each temperature according to the equation, ' - = _. _ + - 5MC+.x XMC+.x 5Mc+.x XMc+5MC+ (0 l“) in which X — and XMC+ sum to one and are relative mole Mc+.x fractions of MC+.X- and MC+ at R = l and the total cesium concentration of 0.01793 M (the highest concentration at which the 1:1 complexation was studied). The mole frac- tions were computed from the 1:1 complex results (Table 33). These average chemical shifts were then used as constants in the calculations. The results are given in Table 39b. The average standard deviation of the chemical shifts im- proved slightly (66 = 2.1 ppm) upon correction for the MC+ concentration, but was still higher than the experimental error. In the case of cesium tetraphenylborate, we could not . fit the data to Equations 0-11 and 0-12 simultaneously. Apparently this is because the dissociation of the 2:1 ion-paired complex is large enough to result in very small activity coefficients which invalidates the use of the 209 Debye-Hfickel equation. In addition we expect both MCE + _ and MC .X to have the same chemical shifts. Therefore, 2 we only fit the data with Equation 0-11. Two procedures were used,similar to the cesium iodide case. First, com- plete association of the 1:1 complex with tetraphenylborate at R=l and (Cs+)t = 0.01 M was assumed. The results are given in Table 00a. Second, an approximate correction for the presence of (MC+) was made in 6 _ according to Mc+.X Equation 0-10. The results are given in Table 00b. In the case of cesium tetraphenylborate a large improvement in 55 occurs when this correction is made. This is a result of weaker association of the cesium tetraphenylborate 1:1 complex compared to the cesium iodide 1:1 complex. The cesium iodide and cesium tetraphenylborate results (Tables 39a and 00a) show that the chemical shift of MC +.X- is essentially the same for both salts. This value 2 also agrees well with the chemical shift of the 2:1 complex in other nonaqueous solvents (201). The value of KX2 for CsBPhu is larger than that for CsI which reflects the dif- ference in ion-association of their corresponding 1:1 and 2:1 complexes. Finally, all of the thermodynamic parameters for the formation of the ion-paired 2:1 complexes are very well determined. Therefore this simple model (Equilibrium 0-12) describes the main features of these systems very well. The only problem is that the average standard devia- tionscfi‘the chemical shifts are higher than the estimated experimental error. However, these high values for 65 209 u + Debye-Huckel equation. In addition we expect both M02 + _ and MC2.X to have the same chemical shifts. Therefore, we only fit the data with Equation 0-11. Two procedures were used,similar to the cesium iodide case. First, com— plete association of the 1:1 complex with tetraphenylborate at R=l and (05+)t = 0.01 M was assumed. The results are given in Table 00a. Second, an approximate correction for the presence of (MC+) was made in 6 _ according to Mc+.x Equation 0—10. The results are given in Table 00b. In the case of cesium tetraphenylborate a large improvement in 66 occurs when this correction is made. This is a result of weaker association of the cesium tetraphenylborate 1:1 complex compared to the cesium iodide 1:1 complex. The cesium iodide and cesium tetraphenylborate results (Tables 39a and 00a) show that the chemical shift of + _ MC .X is essentially the same for both salts. This value 2 also agrees well with the chemical shift of the 2:1 complex in other nonaqueous solvents (201). The value of KX2 for CsBPhu is larger than that for CsI which reflects the dif— ference in ion-association of their corresponding 1:1 and 2:1 complexes. Finally, all of the thermodynamic parameters for the formation of the ion-paired 2:1 complexes are very well determined. Therefore this simple model (Equilibrium 0—12) describes the main features of these systems very well. The only problem is that the average standard devia— tionscfi‘the chemical shifts are higher than the estimated experimental error. However, these high values for 56 210 Table 39. Thermodynamic Parameters for the Formation of the 2:1 Complex of CsI with l8-crown-6 in Methyl- amine at 25.0°C; Assuming Complete Ion-Pair For- mation of Both the MC+ and MC2+ Complexes at (Cs+) = 0.02 M. Assumptions Only the equilibrium MC+.x' + C : MC:.x‘ was considered._ Adjustable Parameters a) Complete Formation b) 5MC+ X‘ Corrected According to + _ . of MC .X at R=l and the Equation (0-10) + (Cs )t = 0.02 M KX2 = 0.66:0.11 0.29:0.09 M‘l AH§2 = -5.09:0.09 -5.00:0.08 kca1.mo1e‘l 6MC2X = -06.02:l.35 -08.6tl.22 ppm 6% = 2.05 2.1 3 ppm 1 (AG298.15)x2=-O'86i0'01 kcal.mole AS;2 = -15.22:0.27 e.u. 211 Table 00. Thermodynamic Parameters for the Formation of the 2:1 Complex of CsBPhu with 18-Crown-6 in Methyl- amine at 25.0°C; Assuming Complete Ion-Pair formation of both the MC+ and MC2+ Complexes at (Cs+) = 0.01 M. Assumptions Only the equilibrium MC+.X‘ + c 2 Adjustable Parameters MC+.X- was considered. a) Complete Formation of MC+.x' at R=1 and (Cs+)t = 0.01 M b) 6MC+ X‘ Corrected According tx>theEquation (0-10) KX2 = 18.3:0.95 27.00i0.5l M“l AH§2 = -7.12i0.27 -6.99iO.ll Kcal.mo1e"l 6MC2X = -5l.2il.07 -09.07:0.00- ppm 55 = 2.7 1.1 ppm (A0398.15)X2=-1.95:0.01 xcal.mole'l A80 = -16.90:0.37 e.u. x2 212 are not surprising since the complete scheme (Equilibria 0—3) was not used but rather only the major equilibrium (0—12) was used to describe both the mole ratio and the tempera- ture dependence of the chemical shifts. In addition to the inexactness of the model, experimental errors can also pro- duce high standard deviation in the calculated chemical shifts. Although the calculations based on this simple model are sufficient for the determination of the thermo- dynamic parameters, attempts were made to refine the cal- culations and to assess the factors which might cause the high average standard deviations in the chemical shifts. These factors are: 1, The neglect of the concentration of (MC+) at the total concentration of the cesium cation is one of these factors as described earlier. An approximate correction to 6 due to the presence of MC+ was made according MC+.X- to Equation (0-10). The results were given in Tables 39b and 00b. The value of 55 decreased substantially in the case of cesium tetraphenylborate but only slightly in the case of cesium iodide. It was mentioned that since the cesium iodide 1:1 complex is more strongly associated than the cesium tetraphenylborate 1:1 complex, the contribution of MC+ in 5 - is less important for the former. In Mc+.x addition, it should be noted that the calculated values for 3 _ are merely approximations. These values were MC+.x not calculated exactly at the total cesium concentrations 213 where the mole ratio studies were carried out, but at the highest concentrations where 1:1 complex formation was studied. Also, plots of 6 - versus temperature were MC+.x curved. Since the 1:1 complex formation and the mole ratio studies were not carried out at the same temperatures, the values of SMC+ - were evaluated approximately from these X curves. However, even with such approximations the de- crease in 56 for cesium tetraphenylborate indicates the importance of the contribution of MC+ to 6MC+.X'; .2. The formation of MC+.X_ from M+.X' and C is not complete nor of the same extent at various temperatures. Therefore the presence of M+.X- would contribute to the average chemical shift of the 1:1 complex. The concentra- tion of M+.X- depends on the concentration of the ligand through, (MC+.x‘) KX(C) (M+.X‘) (0_15) and can be obtained at each concentration and tempera- ture. Then an average chemical shift for MC+.X- can be calculated according to, 1 _. 6 = X 6 MC+.x' MC+.x ' (“‘16) +X MC+.x’ M+.X’6M+.X in which X's are the relative mole fractions of the species 210 _+ _: -+_' - . (XMC+.X XM+.X l), and 6MC .X is defined by Equation (0-10). Equation (0—16) approximately corrects for the contributions of both MC+ and M+.x‘ to 6MC+ x“ The addi— tiOnal correction for (M+.X_) did not change the values of 56 even though the errors in the chemical shifts of some of the points at high temperatures became more random upon this correction. It should be noted that in the calculations of these corrections, the thermodynamic parameters obtained by fitting the concentration dependence of the chemical shift of the 1:1 complexes were used. These parameters, in turn, were obtained from ion association parameters of the salts. Therefore, the accumulation of errors increases the un— certainty of these corrections. 3. The preparation of samples for the mole ratio studies was described in Chapter II. Since we did not use stock solutions of the salts, the total salt concentrations were not exactly the same at various mole ratios, and in the treatment of the data, average values were assigned for the total salt concentrations. To examine whether this ap— proximation affects 56’ the total concentration of the salt at each mole ratio was used as an additional variable to— gether with its proper standard deviation in the treatment of the data. This correction caused noticeable improvement in 66 for both salts. The values obtained for 56 after all corrections mentioned so far were 1.55 and 0.82 ppm for cesium iodide and cesium tetraphenylborate respectively. 215 The thermodynamic parameters for complexation,after all corrections were made,are given in Table 01. The above mentioned factors which affect 56 could be handled quantitatively. There might be other factors which cause high values in 66 but we would not be able to account for them quantitatively. These are: 0. It was mentioned previously that the ion-pair forma- tion constants of the 2:1 complexes could not be obtained from the NMR data since the chemical shifts of MC+ and 2 MC5.X- are expected to be the same. We expect 6MC+ to 2 be independent of temperature, but the chemical shift of MC2.X- might be temperature dependent as is the case for the chemical shifts of M+.X_ and MC+.x‘. If this were the case, the neglect of the temperature dependence of MC:.X- might affect 66' However, an independent measure- ment of KA2 is required to account for this effect. 5. The mole ratio study of the complexation of CsBPhu by 18-crown-6 in liquid ammonia indicates that the solvent molecules and/or the anions interact with the cesium cations in the 2:1 complexes (Section 3). Even though methylamine molecules are larger than ammonia molecules, we cannot reject the possibility of similar interactions in this solvent. If such interaction occurs, then at least two kinds of ion-pairs (ligand separated and solvent separated) may be present in the solution according to the equilibrium 216 Table 01. Thermodynamic Parameters for the Formation of the 2:1 Complexes of Cesium Salts with l8-crown-6 in Methylamine at 25.0°C. Kx2 + + Only the equilibrium MC+.x’ + C MG;.X- was considered. 6MC+ X' was corrected approximately for MC+ and M+.X’ concentrations. Total concentrations of the salts were used as variables at each mole ratio. CsI CsBPhu KX2 0.03:0.05 22.82t0.35 M'1 AH§2 -6.05:0.08 -7.35:0.12 Kca1.mo1e'l GMC; X‘ -06.1710.56 -09.0010.19 ppm 85 1.55 0.82 ppm deg”l _ -l (AG398.15)X2--0.83:0.01 - 1.85:0.01 kcal.mole AS° = —l7.52:0.27 -l8.00i0.00 ’ e.u. x2 217 "K" MC x I [MC2X1' (0-17) 2 in which MCZX and [MCZXJ' are ligand.separated and solvent separated ion—pairs. Although this equilibrium is not con- centration dependent, it might be temperature dependent. Since the degree of association of the 2:1 complexes with iodide and tetraphenylborate anions are different, the equilibrium is also expected to be anion dependent. The neglect of the above equilibrium (if it exists at all) would affect the average standard deviation in the chemical shifts. 6. E. Mei gt gt. (201) studied the complexation of the cesium cation by dicyclohexano-l8-crown-6 in nonaqueous solvents. The variations of the chemical shifts as a func- tion of the mole ratio were indicative of the formation of both 1:1 and 2:1 complexes in pyridine, propylene carbonate, dimethylformamide, and acetonitrile solutions. However, they were not able to obtain the formation constant of the 2:1 complex in these solvents. The reason for such failure appears to be the unusual linear variation of the chemical shift above R = 2. For example, in acetone, at R < l a linear downfield shift occurs as a function of the mole ratio (Kl > 10“). In the range of l < R < 2 a nonlinear downfield shift occurs with increasing mole ratio. Above R = 2 an upfield shift results when the mole ratio increases. This unusual upfield shift seems to be linearly dependent 218 on the mole ratio. The unusual variation of the chemical shift above R = 2 might be due to the dependence of the chemical shift on the concentration of the complexant. This would interfere with the ability of a simple equilib- rium scheme to describe the system. In the case of the complexation of cesium salts by 18-crown-6 in methylamine this effect (if it exists at all) cannot be seen from the mole ratio plots, since the variation of the chemical shift with mole ratio is large and masks minor effects. The discussion about the possible contributions to the average standard deviation of the chemical shifts clearly shows the high sensitivity of the model to minor approxi- mations. However, in spite of these problems, the simple model with appropriate corrections defines the system very well and the calculated thermodynamic parameters of Table 01 can be trusted with a high level of confidence. In addition to the mole ratio studies, the concentra- tion dependence of the 133Cs chemical shift for CST in the presence of a 6.0-fold excess of l8-crown-6 was studied in methylamine at various temperatures. The results are given in Table 02 and shown in Figure 37. The expected chemical shifts at each concentration and temperature were calculated from the parameters in Table 01 and the solid curves shown represent the calculated values. Although the calculated and experimental chemical shifts follow the same trend, they are not identical, probably as a result 219 Table 02. Concentration Dependence of the 133Cs Chemical Shift of CsI in the Presence of a 6.0-Fold Excess of l8-Crown-6 in Methylamine at Various Tempera- tures. _Eobs (ppm) Conc. Temperature, 0C (M) 25.0 -2.5 -10.2 -l6.2 -32.0 0.00511 66.89 06.26 38.70 32.80 9.5 0.01001 59.13 32.92 23.90 16.09 -3.37 0.01302 50.02 25.90 15.78 7.35 —11.75 0.01768 ————— 17.26 9.05 0.70 _;—__ 0.01932 ”7:58 15.86 5.93 -3.95 -19.89 '32.4° C ' l6.2°C ‘l0.2°C “25°C ‘V CI /I 25.0°C I I I__ 8C’00 0.8 LB 24 [CS’] x IO2 (M) 806s (ppm) 0 40- Figure 37. Concentration dependence of the 133Cs chemical shift of 031 in the presence of a 6.0-fold excess of l8-crown-6 in methylamine at various temperatures. 221 of the problems discussed earlier. In summary, the ion-association equilibria of the 2:1 complexes could not be fully understood. However, even with the average error in the calculated chemical shifts of 1.55 ppm for cesium iodide complexes and 0.82 ppm for cesium tetraphenylborate complexes, the fit is sufficient to de- and AH° Attempts to improve X2 X2° the model did not give more information about the system be- termine the values of K cause of the incapability of the NMR technique to separate different kinds of ion—association. The best values for K AH° and the limiting chemical shift of the 2:1 com- X2’ X2 plexes on the basis of the model used are those listed in Table 01. 3. Complexation of Cesium Tetrgphenylborate by l8-Crown-6 in Liquid Ammonia A. Results Cesium—133 chemical shifts of cesium tetraphenylborate in the presence of 18-crown-6 were measured as a function of (l8-crown-6)/(Cs+) mole ratio (R) at fixed Cs+ concentra— tions. The results for (CS+) = 0.001 M are given in Table 03 and illustrated in Figure 38. Another set of experi- ments was carried out with (Cs+) = 0.0075 M. Measurements at R > 1 were not possible below 10°C due to the insolu- bility of the ligand. The data are given in Table 00 and shown in Figure 39. To examine ion-association of 222 Table 03. Mole Ratio Study of 18C6.CsBPhu Complexes in Liquid Ammonia at Various Temperatures; (Cs+) = 0.001 M. 6obs (ppm) Mole Ratio Temperature, °C (1806)/(Cs+) 10.5 0.5 -0.0 —20.9 -33.0 0.000 118.05 120.70 122.09 125.03 127.68 0.25 106.90 108.10 110.00 111.09 ------ 0.095 98.05 98.21 98.52 100.28 101.07 0.60 90.88 90.18 90.18 90.57 95.26 0.70 91.86 90.62 90.15 90.77 ————— 0.90 86.12 80.09 83.08 82.55 83.71 1.00 81.50 79.99 78.36 77.98 77.90 1.20 77.07 75.88 70.09 73.32 73.01 1.39 73.90 71.77 70.30 69.21 69.76 1.59 70.61 68.90 67.66 66.27 67.00 1.80 68.00 66.11 60.72 60.60 65.57 1.99 66.50 60.78 63.07 63.32 63.90 2.22 65.26 63.01 62.08 62.50 63.07 2.38 63.86 62.08 60.76 61.85 62.62 2.575 63.01 61.06 60.80 61.69 62.85 3.00 60.30 58.36 58.05 59.75 60.99 3.32 58.82 57.81 57.58 58.70 60.80 3.97 56.30 55.72 55.60 57.35 60.06 0.97 53.63 50.01 50.01 56.26 58.98 223 ~40 ‘50 ~€K> Swjppm) ‘70 T - 80 40- +90 50' 500 601- 4110 1 ‘ I I 70.. ,l ‘ / l '240 0.4 08 12 -I20 . // | [Cs'] x IO‘ (M) . _ - 1. ________________ _ so / /\© :30 . . // / sot . ’23, . / // (- y IOO ///b \/ no / / to - / ' ’03" $0 // // «of - v 1!, 1 30 1 l 1 I ' I 2 3 4 5 Mole RafioflB-crown-GMCsBpm ) - —-> Figure 38. Cesium-133 chemical shift versus (18-crown—6)/ (CsBPhu) mole ratio and temperature in liquid ammonia; (05*) = 0.001 M. 220 Table 00. Mole Ratio Study of 18C6.CsBPhu Complexes in quuid Ammonia, at Various Temperatures; (Cs+) = 0.0075M. Sobs (ppm) M016 Ratio Temperature, °C (l8C6)/(Cs+) ‘ 10.5 6.0 —25.5 0.00 112.60 110.96 120.08 0.12 105.89 107.99 113.10 0.395 90.38 91.70 96.70 0.55 83.17 80.01 89.22 0.63 79.29 80 77 ' 85.02 0.725 73.71 70.09 78.98 0.83 68.13 68.67 72.62 0.895 66.02 66.81 ‘ 70.61 0.985 62 50 63.16 65.65 1.09 59.13 ————— 62.07* 1.02 53.01 ————— 57.12* 2.19 06.57 06.80* _____ 2.30 05.95 06.26* ----- 2.69 00.17 00.62* _____ * u . Some precipitate in solution. 225 SODS ( ppm ) . A l 4__. 0 LC 20 3.0 Mole Ratio [l8C6 ]/ [CsBph4] Figure 39. Cesium-133 chemical shift versus (18C6)/(CsBPhu) mole ratio in liquid ammonia at various tem- peratures; o l0.5°C, A6.0°C, o -25.5°C; (Cs+)= 0.0075 M. 226 the salt, the chemical shift of cesium tetraphenylborate was measured as a function of concentration at 6.0°C. The results are given in Table 05 and shown in Figure 00. B. Discussion Figure 00 shows that a downfield shift results as the cesium tetraphenylborate concentration decreases. This indicates that the cesium cation interacts more strongly with ammonia molecules than with the tetraphenylborate anion. The degree of interaction of the cesium cation with ammonia is much greater than that with methylamine, since in ammonia the chemical shifts are much more downfield (compare Figures 00 and 15). The concentration dependence of the chemical shift also shows that cesium tetraphenylborate is associated in ammonia. The determination of the association parameters in ammOnia would require an extensive study similar to that in methylamine which was not possible due to time constraints. However it is expected that ion-association in liquid am- monia would be considerably less important than in methyl— amine solutions due to the higher dielectric constant of the former (D = 23 at -33°C). The mole ratio plots show the for- mation of a relatively strong 1:1 complex followed by the formation of a weaker 2:1 complex. The data above R = l were analyzed according to the equilibrium, II II + K MC + C MC+ 2 assuming that the 1:1 complex is completely formed at 227 Table 05. Concentration Dependence of the 133Cs Chemical Shift of CsBPhu in Liquid Ammonia at 6.0°C. Conc. (M) GODS (ppm) 121.00 0.00000 121.56 0.00079 120.32 0.00363 117.05 0.00755 110.96 0.00899 113.96 228 ((2- % ”6- 3 _ “‘33 l20r l l ' '240 0.4 0.8 - (.2 [05*] x (0° (M) Cesium-133 chemical shift versus concentra— tion of CsBPhu in liquid ammonia at 6.0°C. 229 R = l. The chemical shift of the 1:1 complex was fixed at each temperature as the observed chemical shift at R = 1. " and 6 Two parameters, "K C+ were adjusted for each data 2 M set. The results are given In Table 06a. The enthalpy and entropy of formation of the 2:1 complex were obtained from the Van't Hoff equation with the aid of the KINFIT program. In addition,data at various temperatures were analyzed simultaneously, with a linear temperature dependence for 6 The results are given in Table 05b. Even. though MC‘E' separate fits at various temperatures are good, AH° and AS° obtained in this way have high standard deviations. The simultaneous fit has an average standard deviation of the chemical shift of 0.02 ppm which is about two times the experimental error. A comparison of Tables 01 and 06b indicates that the formation constant of the 2:1 complex in liquid ammonia is much larger than in methylamine. If ion association of the 1:1 and 2:1 complexes could be ignored then one would expect the reverse order, because ammonia is a better electron-donating solvent than methylamine and consequently can compete better with complex formation. The limit- ing chemical shift at high values of R in methylamine is almost temperature independent and practically equal to the values in other nonaqueous solvents (201). Surprisingly, the limiting chemical shift in ammonia is about 90 ppm downfield compared to other solvents. This huge downfield 230 Table 06. Complexation Formation Constant, Limiting Chemi- cal Shifts,and Thermodynamic Parameters for the 2:1 Complex of 18C6, CsBPhu in Liquid Ammonia at Various Temperatures, assuming K1 3 104. a. Separate Fit at Each Temperature [Cs+] _ (M) t C:0.5 K2 (611m)2 1 06 0.0075 10.5 395:252 38.0 13.3 0.39 0.001 10.5 1025180 05.08:0.80 0.01 0.001 0.5 17001125 09.17:0.01 0.28 0.001 -0.0 19021182 09.89:0.08 0.35 0.001 -20.9 25661301 53.50:0.06 0.39 0.001 -33.0 3937i669 57.52:0.36 0.37 ("K59298 = 85611000 M“l (99298)2 1 =_0.0:0.7 xcal.mole'l (AH°)2:l = -3.6:0.67 xcal.mole'l (AS°)2:1 = 1.3:3.3 e.u. b.~ Simultaneous Fit at All Temperatures. Reference Tempera- ture = 25.0°C. "K2" = 609200 M-l 0H°)2:l = -0.91:0.28 xca1.mo1e"l GMC; = 00.07:0.65 ppm bMCE = -0.3:0.l ppm.deg_l E6 = 0.02 ppm (AG298)2:1 =-3.88:0.00 kca1.mole (AS°)2:1 =‘3-u5i0095 e.u. shift could be due to the interaction of the cesium cation with both the anion and the solvent, since the chemical shift of the free cation in liquid ammonia is larger than 122 ppm and the chemical shift of the ion—pair with tetra- phenylborate seems to be larger than 100 ppm (Figure 39). The large temperature coefficient of GMCE might also arise from the interaction of the solvent and/or anion with the cesium cation in the 2:1 complex. In any event, a more quantitative description of the formation of the 2:1 complex is not possible since the ion- association parameters of the salt and the 1:1 complex are not known and also the 2:1 complex may form various kinds of ion—pairs which cannot be distinguished from each other by the NMR technique. 0. Summary The thermodynamic parameters for the formation of the 2:1 complex between l8-crown—6 and cesium salts in methyl— amine and ammonia solutions are summarized in Table 07. + _ KX2 + The 2:1 complex formation (MC .X + C 3 MC2 O AHX2 thalpy stabilized but entropy destabilized in both solvents. .X-) is en— The entropy of formation of the 2:1 complex seems to be almost anion independent, but strongly solvent dependent. However, the enthalpy of formation of the 2:1 complex is both anion and solvent dependent. In methylamine solutions, differences in the stability 232 Icoo Houop pew .Uopoospoo no: mm: Ix.+oz a . m mmme Ix +02 mxx o + 1x +02 .Esfipofiafisqo ocu op mcflppooow powmamcm opoz sumo oseo .oapmh oHoE comm pm moanwfihm> mm pom: who; mpamm one no mQOHpmspcoo .mcoHpMLucoocoo Ix.+z paw +0: pom zaopmefixopddm popooppoo mm: Ix.+ozw m mmma . x. o: N o + x. o: I + mxx I + Esfippflfifisvo ocp op wcflppooom commamcm opoz wasp ones .s.o ms.osms.m- os.esss.ma- sm.esmm.sa- mmma HIoaoe.Hsos mm.esad.s- ma.ewmm.s- mo.e.me.o- mmma Huoaos.asos so.owmm.mu Ho.owmm.en Ho.ohmm.ou mxflma.mmmosv mficofie< oceemwmnuoz mafiEmamnpoz Lopoempmm c m c m C no .mcofipzaom aficoEE< paw oCHEmahnpoz :fi muamm Edfimoo pew woma mo ondEoo Hum on» no soapmegom can now mpouoEMme oHEmcmpoELoze .w: magma 233 of the 2:1 complexes of cesium iodide and cesium tetra- phenylborate with 18C6 are mainly determined by the en- thalpy contribution to the free energy of formation. The larger complexation constant for cesium tetraphenylborate compared to that for cesium iodide reflects the difference in the degree of ion—association of their corresponding salts and complexes in methylamine solutions. The much more positive entrOpy of formation in liquid ammonia might be due to the stronger solvation of the cesium cations by small ammonia molecules than by methylamine molecules. However, more data in other nonaqueous solvents are needed to rationalize the thermodynamic parameters of the complexa- tion. 0. Conclusion The mole ratio (R = (l8-Crown-6)/(Cs+)) and tempera- 133 ture dependence of Cs chemical shifts in methylamine and in liquid ammonia show the formation of a relatively strong (K1 3 10”) 1:1 complex followed by formation of a weaker 2:1 complex. If we could neglect ion-association of the salts and the complexes, the data could be analyzed according to the simple equilibria, 230 However, the results obtained from these equilibria were unsatisfactory since the formation constantsand the limiting chemical shifts of the 2:1 complexes were anion and solvent dependent. In addition, the chemical shifts of the 2:1 com— plexes,especially for Csl, were strongly temperature dependent. The chemical shifts at R = 1 were also very different for different anions and solvents. Therefore it is clear that both the 1:1 and the 2:1 complexes are associated in methyl— amine and presumably in liquid ammonia as well. The concentration and temperature dependence of the 1:1 complexes in methylamine showed that three processes are involved in complex formation. These are ion—associa— tion of the salt, ion—association of the 1:1 complex and simple complex formation. A complete analysis of the CsI data (with reasonable assumptions) was carried out and the thermodynamic parameters of the complexation processes were determined by using the ion—association parameters of CsI from Chapter III. The complexation formation constant (M+ + C is MC+) should be independent of the anion. There— Vfore KC and AHg obtained from the Csl data were used in the treatment of the CSBPh0 data and the thermodynamic parameters for the complexation were obtained. The ion-pair formation constants of the 1:1 complexes are of similar magnitude to the ion—pair formation constants of the salts. This could indicate that both the salts and the 1:1 complexes form mainly "loose" ion—pairs and only small relative concentrations 235 of contact ion—pairs are present in methylamine. However, the separation of the effects of these two types of ion—pairs was not possible by NMR techniques. The equilibrium con— stant for the reaction M+.X- + C 58 MC+.X_ shows a trend with, SCN- < l' < BPhu— which indicates a competition be- tween ion—pair formation of the salt and complex forma— K tion. The formation of the 1:1 complex (M+ + C E: MC+) is enthalpy stabilized and entropy destabilized. The thermodynamic parameters for the formation of the 2:1 complex of CsI with l8—crown—6 were obtained from the equilibria, K X2 MCX + C 2 MC X 2 M0: + x‘ K22 MC2X However, since the complex and the ion-paired complex have the same chemical shifts, KA2 is only "determined” from changes in the activity coefficient, which results in a high standard deviation for this parameter. The situa- tion for CsBPhu is even worse since the degree of dis- sociation of the 2:1 complex of this salt is expected to be large and therefore activity coefficients become too small to be estimated by the Debye-Huckel equation. Be— cause of these complications we treated our data according KX2 to the simple equilibrium MCX + C 2 MC2X. In this treatment, 236 weighted average chemical shifts were assigned to MC+.X- in order to correct for the presence of MC+ and M+.X- species. In addition the total concentration of the salt was used as a variable at each mole ratio. The thermodynamic param7_ eters obtained in this way were well—determined. The only deficiency was that the average standard deviation in the calculated chemical shifts, E , was larger than the ex- 6 pected experimental error. Various effects might cause this deficiency such as the interaction of the solvent molecule with the cesium cation in the 2:1 complex, the neglect of dissociation of the ion-paired 2:1 complex, and the neglect of the concentration dependence of the chemical shift. However, even with 36 m 0.8 to 1.5 ppm the system is well defined, and improvement of the model in order to decrease 36 would probably not have an appreciable effect on the calculated thermodynamic parameters. The mole ratio (18-Crown-6)/(CsBPhu) data in liquid ammonia were analyzed according to the equilibrium MCX + C K82 MC2X or MC+ + C K32 M02. The formation constant of the 2:1 complex of CsBPhu with 18—Crown-6 was much larger in liquid ammonia than in methylamine which reflects the difference in the degree of association of the complexes in these two solvents. However, since the formation constant of the 2:1 complex in ammonia is both temperature and concentration dependent, it is clear that the 1:1 and the 2:1 complexes are also associated in this solvent. CHAPTER V 1. RUBIDIUM-87 NMR INVESTIGATION OF RUBIDIUM SALTS AND THEIR COMPLEXES IN AQUEOUS AND NONAQUEOUS SOLVENTS 2. LITHIUM-7 NMR STUDY OF COMPLEXATION OF LITHIUM SALTS BY C211 IN METHYLAMINE AND LIQUID AMMONIA 237 1. Introduction The concentration and counter—ion dependence of the chemical shift of all alkali cations, except the rubidium cation, have been studied extensively in our laboratories (125,126,129—135) and elsewhere (120—122,120) by the alkali metal NMR technique. Similarly, the thermodynamics of com- plexation of all alkali cations but the rubidium cation, has been investigated by this method (152,170,175,176, 200—209). The reason that 87Rb chemical shift studies are sparse is that 87Rb NMR lines are very broad due to the large quadrupole moment and large Sternheimer antishielding factor (210) of this nucleus. For the sake of completeness, we have used 87Rb NMR in an attempt to investigate ion— association and complexation of rubidium salts with 18— crown—6 and C222 in solution. Three solvents were used in this study: water, methanol (protic) and propylene car— bonate (aprotic). Landers' (8) suggestion that the exchange rate of the lithium cation between the solvated and C211—cryptated species in methylamine is extremely slow, motivated us to study this system by 7Li NMR technique. In addition a mole ratio study of the complexation of the lithium cation by C211 in liquid ammonia was carried out. The results of these experiments are described in this chapter. 238 2. Investigation of Rubidium Salts and Their Complexes in Aqueous and Nonaqueous Solvents A. Salt Solutions The concentration dependence of the 87Rb chemical shifts and linewidth of rubidium salts in water, methanol, and propylene carbonate (PC) was studied. The results are given in Table 08 and Figures 01 and 02. Our results for aqueous rubidium bromide solutions in the range of 0.2 to 1.0 molar agree well with those of Deverell and Rich- ards (121). In addition, pulsed Fourier Transform tech— nique allowed us to study solutions as dilute as 0.02 M. Over the complete concentration range, the variation of the chemical shift with concentration is nonlinear. However, a plot of chemical shift versus the mean molar activity of the solution is linear and in agreement with the results obtained by Deverell and Richards. The same behavior has been observed for aqueous cesium bromide and iodide solu— tions in our laboratories (211). As has been pointed out in Chapter III, the origin of this behavior is not known quantitatively. In methanol (D = 32.7) and propylene carbonate (D = 69.0) solutions, the curvature in the plot of chemical shift versus concentration is more pronounced at lower concentrations. The mean activity curves show linearity at high concentrations, but become nonlinear as the concentration decreases. In these two solvents, 200 00 000000 sec 00.0 c .pco>aom 65p mo zpflafioflpdoomzm 00062005006 6:0 mm 0062 000006 onoaomH 050 00% 660060000 600 mpmficm 0000Eosom 000 00.0 000.0 000 00.0 000.0 000 00.0 000.0 000 00.0 000.0 000 00.0 000.0 000 0.00- 000.0 000 0.001 000.0 000 00.0 000.0 000 0.00- 000.0 000 0.00. 000.0 000 00.0 000.0 000 0.00- 00.0 000 0.00. 000 0 000 00.0 000.0 000 0.00- 00.0 000 0.00- 000 0 000 00.0 000.0 000 0.001 00.0 000 0.001 000.0 000 00.0 AS000.0 000 0.001 00.0 000 0.00. 000000.0 000 00.0 :00000.0 000 0.00- 0000.0 000 0.00- 000.0 I--- 0.0 0.0 0000 Asddv 020 0000 05000 00: 0000 02000 020 m\0>< 0900 .0200 m\0>< mno .ocoo m\0>< .mnomV .ocoo 000 000 000 000 opmcooLMo 0c00>dopd Hocmnpoe 0mm :0 000 :0 0000 :0 0000 .moLSmedeoB pcofinE< pm muco>aom moses c0 h0000 20000000 0o hsse0sos00 0:0 000000000 00002000 00120000030 .00 o0ome 201 .pqm>Hom oEmm ogp CH mpCHOQ hogpo mo page mo moEHp 03p pmmoa pm 606 m AppflonHH can mpgflgw 00005050 CH mcoflpmfl>mp pmeCMpm .mamcwflm 0m00: ogp 0o omzmoom p .00 Q0 mm 000 pew .HOQNonE 20 N: omfl .00p63 Q0 mm 000 CHQpfls ow ©006> osw Am\0><0 pgmflocImHQQ pm msppflsocfiqo 0cm . .00 Q0 sad 00.00 0ocw£poe Q0 Egg m.00 .ONE :0 Egg 00.00 :0g003 op 00605006 006 mpwflnn HMOHECQQQ .poszflpcoo I monoCQoom .0oss0eso0 .00 00000 242 E 2 052 0.64 O.L06 ' 0: 3% 3L d '8 6CD ‘1' . \ \ 55- \\ ‘ \ £5 _L. J_ 41, L \’ O 0.2 0.4 0.6 0.8 LO ‘ Conc. (M) Figure “1. Rubidium-87 chemical shifts of rubidium bromide versus concentration (-) or mean molar activity (---) in aqueous solutions. 2M3 0.02 ' 0.04 0.06 -28- - ’g -26 - CL .9 g -24 060 -22 .- -20 _ l l I l L 0.02 0.04 0.06 0.08 DJ 0 Conc.(M) Figure 42. Rubidium—87 chemical shifts of rubidium bromide in methanol_(—) or rubidium iodide in propylene— carbonate (-——) versus concentration (0) or mean molar activity (A ) of the solution. 244 especially in methanol, ion—pair formation is also res- ponsible for chemical shift variations. Attempts were made to fit the data to a simple ion—pair equilibrium in these two solvents. The variation of chemical shift with concentration does not follow the simple ion—pair model, presumably because other interactions in solution are also responsible for chemical shift variations. B. Complexation Attempts were made to study the complexation of the rubidium cation by lB—crown—6 and cryptand—222 by using 87Rb NMR techniques. The results in H20, methanol, and PC are given in Table A9. The exchange rate of the rubidium cation between the solvated and Rb+-18—crown—6 complexed species in aqueous solution is fast on the NMR time scale (213). Therefore, a single population averaged signal is expected for this system at all mole ratios. At mole ratio (lB—crown-6)/ (Rb+) = 0.A7, where the concentration of RbI is 0.1A m, the linewidth of the signal is about 1100 Hz; therefore the chemical shift determination is not precise. If there is any variation in the chemical shift as a function of the mole ratio, this change is not detectable because of line broadening, and a mole ratio study of this system is not possible. Above mole ratio 1, the signal is undetectable at this concentration. The complexation constant of 245 0 00.0 00 00.0 0000 0 000.00- 00.0 000 0.000.00- 00 00.0 00 0000 00.0 000 0 00.0 000 0.000.00I 00000.0 000 0.000.00I 00.0 0 0mm 0.000.00I 00 00.0 00 00 0000 00.0 0000 0 m0.0 0 000 0.00m.m 00 000.0 0 0 0000 000.0 000 0. 00.0 000 0.00 0.0 00.0 0 000 00.0000.0 00 00.0 0 0 0000 00.0 000 0000 00 0 00.0 0000 00 m- 00.0 0000 00 0- 00.0 0000 00 0 00.0 0 000 0.000.: 00 00.0 0 0 0lczopel00 00.0 000 0 0.0 00 0.0 0000 0.00 0.0 00.0 0 000 00.0000.0 00 00.0 0 0 0103000100 00.0 000 IiIrIIIIrIrIrIIIIIII1IIIIIIaI1IlIItIIIIIIIltlrlltltrlllltllrlllllII 0000 00000 0000 00000 000>0om 000000 020 0000 m\0>< 0900 .9000. 0002 .2000 0000 I1IIlJ1IIIIIIIIII1IIIrIIIIIJIIIIIIIIIIIIIIIIIIIIIItIlllllllllllltllllllllll rIIIIJ1IIIIIIIIIIIIIIIIIII|IIIrIIII .00000 0002 0+000\00c00000 msmp0>000000 00000000 00-00000000 .00 00000 246 .x000 00050 >00>0 .0050: 00 00 00>00000 003 002000 020 .m . . 0m 0 op 00 0 000 0.000.001 00.0 000 0.000.001 00 00.0 00 0000 00.0 000 1111111111111111111111111111111 11 mwmv Asgav 000V 0000m pcm>0om 0Q0000 sz 0000 >< 0000 .QEOB 000: .0500 0000 HHHw11111111111111111111111111111 .Umscflpsoo .m: manmfi 247 Rb+-18-crown-6 in water is 36.3 (173), so that at a mole ratio of n.5, more than 90% of the Rb+ should be complexed. Therefore, above mole ratio 1, the signal disappears because the line is very broad and weak, and cannot be discriminated from the noise. However, a broad signal (Avl/2 >3000 Hz) was detected above a mole ratio of one for a solution which was 0.5 M in rubidium iodide. The rate constants for the dissociation of Rb+C222 cryptate in H20, methanol, and PC are 1M0 (21M), 0.8 (215), and 0.17 s‘1 (216), respectively. Therefore, two NMR signals are expected at O < mole ratio < l in all three solvents. Dye et al. (217) reported a peak for aqueous Rb+C222 which is about 50 ppm downfield from that of 0.1 M aqueous rubidium iodide solution with a line- width of 1300 Hz. However, they were not able to observe the complexed peak in alkali metal solutions containing cryptand 222. The authors also reported a broad signal (Avl/2 2 H000 Hz) for the complex in methanol solutions at about 100 ppm downfield from a 0.4 molar rubidium iodide solution in methanol. Our attempts to observe similar signals failed even though different experimental condi- tions, such as different delay times, different sweep .widths, different concentrations, and different spectrom- eters were used. The complexation formation constant of Rb+-C222 in H u 20, MeOH, and PC are of the order of 2 x 10 , 9.5 x lO8, and l x 109 (216) respectively. In methanol we have been able to see a very small peak 248 at mole ratio of 0.98 even though the fraction of the free cation is very small. A broad signal (Avl/2 = 1300 Hz) was also observed in PC at mole ratio 0.U6. The above facts show that if the linewidth of the complexed cation signal in water is indeed only 1300 Hz we should have been able to observe it easily. In the case of Rb+c222 in water no peak was observed within i200 ppm of the free rubidium cation signal. It appears that the complexed signals broaden as a result of some kind of exchange. Perhaps the exchange occurs between inclusive and exclusive complexes as in the case of the cesium cation complexes with 0222 (152). Another possible explanation is that the true line- width of Rb+-C222 in water is much larger than that re— ported (217). However, we expect narrower lines for this complex than for the complex with 18-crown-6. 3. Complexation of the Lithium Cation by C211 in Methyl- amine Solutions of lithium bromide and C211 in methylamine were prepared as described in Chapter II. The solutions were mixed and the 7Li NMR spectra of the resulting solu- tion were taken more than 90 times over a 2.5 hour period at -51°C. Only part of the data is given in Table 50, since no distinct changes occur between the reported data points. All data consisted of two signals, one for the free cation and the other for the complexed cation. The 249 Table 50. Lithium—7 Chemical Shiftsrofthe Free, 6F, and C211—Cryptated, 5 Lithium Cation in Methylamine as a Function of Time. (211)/(Li+) = 0.5, (Li+) = 0.069 M. Time 5F Sc Hr Min Sec (ppm) (ppm) 0 (mixing) 15 0.75 —O.82 40 0.66 -0.93 4 02 0.62 ~l.03 7 19 0.55 —l.05 ll 24 0.59 -l.06 13 43 0.70 —0.90 20 26 0.76 -0.82 23 25 0.84 —0.75 27 40 0.84 —0.79 35 00 1.00 —0.65 46 56 1.03 -0-52 57 14 1.05' -0.52 l 04 00 1.08 —0.52 1 ll 00 1.04 —0.58 l 38 59 0.97 -o.61 2 22 37 0.96 —O.56 96 1.16 -o.u2* >10 days 0.67 —O.98* * Precipitate was formed in the solution. 250 chemical shift of the reference solution, which was measured both before and after the experiment, had values of 15.11 ppm and 15.19 ppm respectively. It was noted that some precipitation occurred in the solution after 2—1/2 hours even though both C211 and lithium bromide were completely soluble at the time of mixing. This precipitate could not be dissolved even after 96 hours at room temperature. The precipitate dissolved gradually in a few days but the chemical shifts of the free and complexed species were 0.67 and —0.98 respectively; It should be noted that the difference in the chemical shifts of the free and complexed lithium cation was constant at all times (1.59 i 0.04 ppm). The ratio of the intensities of the two signals remained almost constant at a value of 0.8 — 0.9 during the 2-1/2 hour period. The signals had almost equal intensities after 96 hours. The final chemical shift of the complex was -0.42 ppm, in good agreement with the corresponding chemical shift in other solvents (170). In addition, the chemical shift of lithium bromide in methylamine was concentration independent over the concentration range studied (0.0058 to 0.094 M), and had a value of 1.11 ppm at —52°C. The data show that the exchange rate of the lithium cation between the free and bound species is slow on the NMR time scale, but it is not extremely slow, since most of the complex is formed immediately after mixing (within 251 15 seconds). The data also demonstrate that the chemical shifts of both species change with time. The fact that the difference between the two chemical shifts is constant, and the chemical shifts of the dissolved solution after a few days is equal to the chemical shifts at the time of mixing, suggests that precipitate formation is involved in the chemical shift changes. It is possible that the C211 sample had absorbed small amounts of carbon dioxide during weighing. Then if the formation of the complex occurs Via the formation of an exclusive complex, it is probable that both the lithium cation and lithium exclusive complex form ion—pairs with the carbonate or bicarbonate ions which would affect the chemical shift. The role of the precipitate cannot be assessed without information about its composi— tion. It is important to point out that the first attempt to make the solution failed, because even before the solu— tions were mixed, precipitation occurred in the lithium bromide solution. In this attempt, some of the isopropanol— dry ice solution, which had been used for cooling, col- lected on the Kontes Teflon vacuum valve accidentally. The isopropanol—dry ice solution may then have been trans— ferred to the lithium bromide solution to form carbonate in the solution. In any event, our experiment showed that the complexa- tion of the lithium cation by C211 is not as slow as has been predicted (8)- 252 4. Complexation of the Lithium Cation by C211 in Liquid Ammonia A mole ratio study of the complexation of lithium bromide solutions by cryptand 211 in liquid ammonia was carried out. The results are given in Table 51. The ex- change rate of the lithium cation between the solvated and complexed species in liquid ammonia (as in other sol- vents (170)) is slow on the NMR time scale, and therefore two signals were observed below mole ratio 1. The chemical shifts of the free and complexed species are nearly tem- perature independent, and the complexed signal is 1.9 ppm upfield from the solvated Li+ signal. The chemical shift of the complexed lithium cation agrees well with the results in other nonaqueous solvents (170) indicating that the lithium cation is effectively isolated from the solvent and the anion by the ligand. 5. Conclusion Rubidium-87 chemical shifts of rubidium salts in water, methanol, and propylene carbonate show a nonlinear de- pendence on concentration over the complete range of con- centration which was studied. However, the chemical shifts in water are linearly dependent on the mean activity of the solution. In methanol and propylene-carbonate, the non- linearity in the plot of the chemical shift versus 253 me.o- m:.o- »-- om.o- -- s:.o- -- 33.0- -- OH.H :m.o- Hm.H wm.o- om.H mww.o 52.0- mm.a mm.o- m:.H m:.o- w:.a mwm.o wm.o- mm.H mm.o- mm.a sm.o- :m.a mm:.o mm.o- om.fl 53.0- s:.H sm.o- em.H mwm.o -- H:.H -- m:.H -- om.e 00.0 poonQEoowl momma UmonQEoom momma poonQEoow oopmw oepmm mama coo.em- ooo.w ooo.m AEQQV on© .2 No.0 n A+eqv .eflccee< cussed he Hem cemeeseo eo cocchcee see an emeu eo eehem HmcHEceo flue cap eo sesem checm cue: .Hm cease fl 254 concentration cannot be accounted for only by considering ion-pair formation in the solution. This suggests that other effects are also responsible for the chemical shift variation. Rubidium-87 NMR signals are broad in the presence of C222 and especially l8-crown-6 in aqueous and nonaqueous 87Rb NMR techniques are not suitable solvents. Therefore for studies of the complexation of rubidium salts by crown ethers and cryptands in solution. Lithium-7 NMR data show that the exchange rate of the lithium cation between free and C211 cryptated complex species is slow on the NMR time scale in methylamine and liquid ammonia. The chemical shift of the complexed cation is the same as in other solvents which indicates that the lithium cation is effectively isolated from the solvent and the anion in the complex. CHEERVI SUMMARY AND SUGGESTIONS FOR FURTHER STUDIES CHAPTER VI SUMMARY AND SUGGESTIONS FOR FURTHER STUDIES 255 1. Summary The concentration and temperature dependence of 133CS chemical shifts of cesium salts in methylamine was studied. The variation of the chemical shifts of cesium iodide and cesium tetraphenylborate as a function of concentration indicated strong ion-association of these salts in methyl- amine. However, the data did not follow the behavior ex- pected for simple ion-pair formation. Instead of leveling off at high concentrations, the chemical shift versus con— centration showed a gradual downfield shift for CsI and a gradual upfield shift for CsBPhu. The chemical shift of CsSCN was concentration independent in the range of concen- tration and temperature studied. The cesium iodide and cesium tetraphenylborate data at various temperatures were analyzed according to the equilibria for the formation of ion-pairs and two kinds of triple ions (cationic and anionic). In this analysis, determination of all the thermodynamic parameters from the NMR data was not possible. Therefore, we had to make justifiable assumptions about some of the parameters. For example, the triple ion formation constants were cal- culated from the Fuoss equation and used as constants; linear temperature dependence was considered for the chemical shifts of the free and ion-paired cations; equal 256 257 probability for the formation of the two kinds of triple- ions was considered; and the chemical shift of the cationic triple—ion was assumed to be the same as the chemical shift of the ion-pair. A long extrapolation was required to ob- tain the chemical shift of the free cation, and since the fraction of this species was small even at the lowest concentrations, the determination of this parameter from just the data with CsI and CsBPhu was not possible. The fact that the chemical shift of CsSCN was inde- pendent of concentration, together with other experimental facts led us to propose that the chemical shifts of the free cesium cation and the ion-paired CsSCN are the same. This was the major assumption in the treatment of the data. Other assumptions did not have significant effects on the ion-association parameters. The ion-pair formation constants, K at 25.0°C for ip’ these salts (averaged from the models with different as- sumptions) were found to be: (Kip)CsI = (2.65:0.19) 5 -1 O O = + -l X lo M Wlth (AHip)CSI 3.7—O.3 kcal.mole , (Kip)CSBPhu 4 -1 ° 0 _ M Wlth (AHip)CsBPhu - u.0:1.o kcal. mole-l. The equivalent conductance of Csl in methylamine = (1.30:0.19) X 10 was measured as a function of concentration at -15.7°C. The conductance data were analyzed according to various conductance theories. The value of (Kip)CsI obtained from conductance at -15.7°C depended on the theory ap- plied and had values ranging from 8.7 x 103 to 1-4 x 106 M-l, 258 The Onsager limiting law fit the data the best but Kip obtained from the limiting low was only about 35% that ob- tained from NMR data (corrected to -15.7°C). Cesium-133 chemical shifts were measured as a function of the (l8-Crown—6)/(Cs+) mole ratio (R) and temperature. The variation of the chemical shift with R indicated the formation of a relatively strong 1:1 complex followed by the formation of a weaker 2:1 complex. The formation constant of the 1:1 complex is too large to be calculated from mole ratio studies at a fixed total cesium salt con- centration. The data for CsI and CsBPhu above R = l were analyzed at various temperatures according to the Kc2 + Kx2 equilibrium MC+ + c 1 M02 or MC+.X‘ + c z MC+ ' 2.x The fact that KC was anion dependent together with the 2 behavior of simple salts in methylamine and with the fact that the limiting chemical shift of the 2:1 complex was temperature dependent, indicated that the 1:1 and presum- ably also the 2:1 complexes are associated in methylamine. Therefore, to analyze the data above H = 1 it was necessary to investigate the ion-association of the 1:1 complexes. The concentration and temperature dependence of the chemical shift of the 1:1 complex showed that ion-pair formation of the salt competes with complex formation and that the 1:1 complex also forms ion-pairs. The variation of the chemical shift as a function of the concentration of the 1:1 complex and temperature was analyzed according 259 to the equilibria, + _ M .x M+ + + Kt M+ + K K _ i _ - t _ _ M+ + X 2p M+.X + x z x .M+.X + + c 0 ii KC ,p, KX K + _ A _ MC + x z MC+ X The internally consistent thermodynamic parameters ob- tained from NMR studies of ion-association of the salts in the absence of complexant were introduced as known constants in the above equilibria. The KA values for CsI and CsBPhu ((1.51ro.06) x 105 and (1.16:0.3u) x 10 M.1 at 25.0°C respectively) proved to be comparable to 4 the K values for the uncomplexed salts, indicating that 1p the formation of solvent-separated ion-pairs of both the salts and the 1:1 complexes probably dominates over contact- pair formation. The value of Kc’ which is anion independent, 260 was (1.07:0.08) X 10“ M-1 at 25.0°C with AH: = -l6.72i0.08 -l Kcal.mole . Other parameters were found to be. (Kx)CsSCN = (4.87:0.53) x 10"3 < (KX) (6.33:0.40) x 103 < (KX)CSBPhu = responding AH; values of -l3.50t0.73, —l6.40i0.53, and CsI = (8.u:1.u) x 103 m‘1 at 25.0°C with the cor— -l8.8i0.95 Kcal.mole“l respectively. The order in KX for various salts reflects the difference in the degree of ion- association of the corresponding salts in methylamine. The mole ratio data in methylamine for R > 1 were ana- K lyzed according to the equilibrium MC+.X_ + C $2 MC 0 AHx2 with proper corrections applied for the ion-association of + 2.x the salts and of the 1:1 complexes. The thermodynamic parameters were found to be: (K = 4.03:0.05, and l x2)CsI at 25.0°C with (AH O ) _ x2 CsI = -7.35:0.l2 Kcal.mole. The (K = 22.82i0.35 M- X2)CSBPhu O -6.05:0.08, and (AHx2)CsBPhu dissociation of MC:.X- was not included because the ion—pair formation constant of the 2:1 complex could not be obtained from the NMR data, (presumably because the chemical shifts + _ of M02.X and MC+ are nearly the same). The corresponding 2 approximate values in ammonia are: (K = 649144 M-1 l x2)CsBPhu at 25.0°C with (AH = -4.9li0.28 Kcal.mole- ° ) x2 CsBPhu Rubidium—87 chemical shifts of rubidium salts were measured in water, methanol, and propylene carbonate. The variation of the chemical shifts as a function of the concentration was nonlinear in all of the solvents. However, a plot of the chemical shift versus the mean molar activity of the salt was linear in aqueous solutions. 261 The variation of the chemical shift with concentration in methanol and in propylene carbonate did not follow a simple ion-pair formation model. Attempts to study complexation of rubidium salts by l8-crown-6 and cryptand-222 failed either because the NMR lines were broad or the chemical shift changes were small. The exchange rate of the lithium cation between the free and the 211-cryptated complex species in methylamine and liquid ammonia is slow on NMR time scale, consequently (0211) (1.1+) was formed in less than 15 seconds in methylamine. The two signals were observed at 0 < < 1. The complex chemical shift of the complexed species was the same in both solvents and equal to the value obtained previously in our laboratories indicating that the lithium cation is effectively isolated from the solvent and the anion in the complex. 2. Suggestions for Further Studies The studies already made stimulate the following sug- gestions for further studies: (1) It has been shown that the 13305 chemical shift in methylamine and liquid ammonia changes gradually at high concentrations instead of leveling off. The oc- currence of similar behavior in methanol and ethanol solu- tions and with 87Rb chemical shift in methanol and PC solutions suggests that the chemical shift is concentra— tion dependent even when ion aggregates do not form. To study ion—association by the NMR technique accurately, it is necessary to separate the contributions of concentration and ion—association to the chemical shift. Determination of ion-association constants by various methods such as elec- trical conductance measurements, calorimetry, and UV spec- troscopy and the comparison of these results with the NMR results would provide a useful probe for the study of this problem. (2) Since ion—pair formation constants and the limit— ing equivalent conductances, A0, obtained from electrical conductance measurements depend on the theory used, an independent measurement of A0 would help to test the validity of various conductance theories. The best method for ob— taining A0 is to measure both conductances and transference numbers. (3) An extensive study of ion—association and complex formation of cesium salts in liquid ammonia and the com— parison of the results with those obtained with methylamine would provide valuable information about the role of the solvent in these processes. (4) The study of complexation of the lithium cation by cryptand—211 in methylamine has already shown that the exchange rate of the cation between the free and complexed species is slow on NMR time scale, but the reaction takes 263 place in less than 15 seconds. In this study complications arose because of the formation of precipitate in the solu- tion. A more careful experiment is required to explain the complications we encountered. iMore study is also re- quired to understand the mechanism of complex formation. APPENDICES 264 APPENDIX 1 DETERMINATION OF ION-ASSOCIATION PARAMETERS BY NMR TECHNIQUES; DESCRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN A. Simple Ion-Pair Formation The equilibrium for ion-pair formation can be expressed as Kip M+ + x’ + M+.x‘ (lA-l) + - K := (M .XI) = l-d (lA-2) . + _ 1p (M )(x >yi oa2y§ in which M+, X', M+.X- are the solvated cation, solvated anion, and ion-pair species, respectively, the terms in parentheses are the molar concentrations of the correspond- ing species, Kip is the thermodynamic ion-pair formation constant, a is the degree of dissociation of the ion-pair, and y: is the mean molar activity coefficient of the solu— tion. The latter can be calculated from the Debye-Huckel equation, 265 266 -4.l9764xlO6IZ+Z_I/I Y = exp( /_ ) (IA-3) (DT)3/2[l + 50'29 3 1] H- in which Z+,Z_ are the charges of the ions, D is the dielec- tric constant of the solvent, T is the temperature in °K, 3 is the distance of closest approach of the ions in Angstroms,and I is the molar ionic strength of the solu— tion (I = 1/2 EC dZi). i i The observed chemical shift is a population averaged chemical shift and can be expressed as, 6obs = XM+6M+ + XM+.X-6M+.X- = 6M+ + (l'a)5M+.X- (lA’u) where XM+ and XM+.X‘ are the relative mole fractions of the free and ion-paired species, respectively, and 5M+ and 6M+.X- are the chemical shifts of the corresponding species. Three parameters should be obtained for each salt from the fitting of equations lA-2,3, and 4 to the NMR data. These Q. . parameters are Kip’ 5M+’ and 6M+.X-' since 6M+ in a glven solvent is independent of the anion, the above equations can be fit to n_data sets at a given temperature to obtain 3n—1 parameters. In our case, data for CsI and CsBPhu at 25.0°C were used to obtain five parameters. These are, 267 U - 505+.1‘ U = (Kip)CsBPhu U(2) = (Kip)CsI U<5> = 6Cs+.BPhu' U(3) = 5CS+ The values of D, T and 8 were introduced as follows, l/2 Const(JDAT,l) (DT) O const(JDAT,2) 3 = 5.3 A (Chapter III) The value for D was taken from Reference 194. The sub- routine EQN is given on the next page. B. Ion-PairszmxiAnionic Triple-Ion Formation The equilibria for the formation of ion-pairs and anionic triple-ions can be expressed as (lB-l) - - _ + _ M .X + X 2: X .M .X where * ... NMR TECHNIQUE FROM 268 MULTIPLE DATA SET *iitttiittttttt*****i*titiitiitii*tittttitttkiitttttitt FORMATION f“ H DA CH-PAI Y ittiiiittititii*ttiifi*tt‘tiiiifii*tititii‘kiii‘kifiiiiititi’ l i i O 4 TI. F 0N II LoK E0: A64 2LT $9U99:Do Crab-LT LNFAC x...” CANAFr-nwfivlohkoruw PJFHRHFMLLT 9F. RdQNB AKTI: QFKDn FUN. ’9 T DEIEO DCRNKYGTN PRONODLRI ALHAHN.‘ :AHK BPLBPQ NODD ' \l QC )1 1‘ 2Y9! ' .U)’ )0 0.00 0.2)r3p3 ((nUl“ XTllD-VI XC‘YY QCYTY ’v 0x ’ o O’R) 0’0 1.0 31nu)5 0.23041 49(5T I‘VUTI‘P vrnCSMunU T‘XIL XHQF 9M“ C x 0. 9’5) X)\IPJIO 9002 95 T209()l\ .P‘SLHUTI GU(AAC¢F. d9UV‘O H .)FP0TIM ITYo 79809910911912.13) “350 9N01 9 IKOC)!5(5 //41uh-u.\ll\x ’ TT(l\n¢6xxa. DNXO(1RXQ EUU 9 QSCI F0 91 93 NO ’PYTRON 9T0 Q) Q IKXYQSFPO)95)62E01 T 9TJN//IO)(01(U6625 U U UNKXJCNNQVOOSRVI\ C C C \I C C E C C 9 SC C U MC C T CC C A RC C 1.». R TC C H E SC C T P P 5C C IA 8 M NC C S 84.: AC C CN CHT C C 0 pt NC C TF1 FBT IC C SOT OSN C C C A CA EC C TC T T CC C TFN NFS NC C NFOAS AON AC C 01 TC T 0 TC CM IHTS STC SC CC TSFNE NF IC CI A 10E OIC DC CT TRLHCR CHI C CA NTAS F SR NC CI ENC N N T CC CC IEILOFNOLC IC CC CCMAIOOIAE C CSFINECT ITCLQ...C CSTFOHIATTAIE/ C CIGFCCMMFUHMIINC CONE ERILREDtOC C EOTDHOHIOH‘tIC CFRCLECFSDFC: :C COT AV ) )C C SYSRGRLERGI 2C CE T FNIATINQ 0C CECILSIACIAIT TC CPIVABTPINPTA AC C.U.NITOI M1... ...-.0 DC CEOT0:MNEFNVJ JCECCLQ C..UI»CTSIOHNOI( (CU N: : : :NNNNNC: :A: : LICIILS OCNRJCCONSCT‘TOIEERKRIPJRCAN:)) : : : : :T TCI: 0.1 )/(2o*U(2)*XX(1)*(€ANA**?)) )*(GAMA**2) 0910011JDAT olUlM‘ ZR SCNI‘N : C 2 ) no 7R 3L, ( 0T TR Pu OS 6* ) ”Uh-V O 0 AUG .4 + OlXN.L/O .XELT 09 T61 RANSR4 RVU999MVL(3$ISTTPPAAUTUTUCfiEA12))) ))S SCTATTMAHE:A. BMOXMYMMVHP‘NIN 127g Q C; C C C C C C C CCCCCCCCCCCCCCCC C C C2 NAAVUTNTNTCPRN((125 45W NCNM NG(PRG(: DUONXQYOCICLOXOOOTTOOEOEOECLTAXX((( ((0 OCOAOORFLTRFA RSC9OLYCCUFZCDCGCIJNNRCRCFCASGXXUUU UUC CCCGGCAIASAIA l Cc 1C 4 * * ifi*tt***kt~kfii*~kiiii NMR TECHNIQUE **r**ttt**t**tii***ii SET-CONTINUED 2'69 MULTIPLE DATA ION-PAIR FORMATION FROM iii-*ii'tti'i'iii*‘k‘kiti’t‘ktttttiittttti'ii' ************i*‘kitt‘k‘ki‘ki-t‘k-kti'ii‘tik‘k * it ) ) ) \I 1 \I l ’ \I 9 1 ... mi 2 t ( K ... ( K S N t S N T 0 A T 0 S C M S C N t A N t O o G U a C 2 ( C 2 / ( .w / ( \I / ) ) / T A \I \I 1 \I \l A Ml. G \l l\ G \I D A R ) X R ) Iu G A 1 VA A 5 9 9 I\ l\ \l * I.\ l\ A A T U 2 \l T U M H R _ ... 4- R _ A Dr Cu ) i ( Aw. \l mu L S ) 3 A U S T 3 9 A i U 0 I\ M t .x .--d 1 l\ A Q. \I E C U A o \I E C U H- J 2 N 0 I\ G 2 2 N no /\ D- l 9 A l) * 5 (71\ \I O. A l) t r? L) (\I ..L 1 M O \I 3 i~Ql F0 2 M D \1 .3 A5 X4. 2 ( A 0 o 7.) ) ) TR ( A 0 o 7) 91 X o S .6 T2 3:- O 10) 3A 5 6 T2 3F 0 0’1 10 O T. / ... + T (TM (\ Tr / ... + T c1 9 T1; T- QV ) 0* 0 o X G OT S \I 0* O o C(X AE N )A GW T1 0 XORA TR N )A va T1 0 X5 03 0 30 OM \IE) I\ G *IUAH Pu—DO D” )E) l\ G OVA 0. IV V G *C DA lNlOT ))(P OStC DA INIOT )T)4 )X it ’5 0A(GR ) AUTL G*t* /G 0A(GR ) 2 4. 35 ) ’01 A . OMX Q l ( oDnA \I )9 A - CMX Q T. (030 C 9 1 12 Ann 0AX)S . non-0* 0312 Ann DAYTIS . X611 15 _ 9. (...-TI“ oCi0+xl¢ .x .S) 0.9. I‘LL)“ orUtO+slo VA) Orr. 9T. 0 lnu D-NTEO( o cit-L orL+l 0020 PNT-LO‘ o 95E .3E3 E1 E (5 XAAN.*K11(NC 4L-( 04(5 XAAN.*K11(NC CoP9 5P9 N Sl\ EMRA,T\IN. . U 0L + fixutuuLLrbS‘ EM...HAT)N. .U o-L |:LAX EAX o TI\C:A(M5520 0(+HA r10 _XEL7T(C__A(MhGa-O 0(+HA EANTFOFL .L 0T5 E r... EH E F. 2.. S+CHGSA0(CE()TC U1M(:R-9S+CWGSA.(CE()TC UC.J(U UIJ( U U UT U U U NGtE‘BGTU*L()E:NN:G: TGINOiE‘BGTU...L()rr_:.NA:D-(TNNN:(TNNNNNNENAHNNAK 0.5N(A:A:co:)M)RI RANSRQO.BN(A:A:oo:)M)RIDAEAIRIDEARIRIRIMRIRTRIR CIEA::AFKAFCIIZUTMAHE:A.CIBA::ARKAFCIIQUTILTVTUTITMUTUTUTIUTUTUTU __ : .....P-TT-MuI‘N: (L(((TNG(PRG( __ __ : __M-..TITM.I\.N__ (Ll‘((TNS/KIRHNTNSIRTNTNTN‘TNTNTNT DDCDAAAHAF.O FAXFXEOSHFLTHHFABCDAAAAFO EuAXE.-XEOEFROHUFEOEROEOEOEOFPZOEOPLOCL BCDG—nRFUIFVFICXIXRCAIAAbnkTAI-HD._DCRUGRRPOICFICXIXRC..KIHW.FCRPCDAHFRCRACRCIDAACQHCRCQ TL 1 1 5 Q- ..U 7 ‘3 .3 Q. 5 ...-... O ... . O I) CC «C n.» 2 1 . ... T. (C 11 \i 270 *izi’iiiflkitiiii*t******i*‘k*iiiiiiflkii******t***i**itt*i'*it * ION-PAIR FORMATION FROM NMR TECHNIQUE * MULTIELE DATA SET-CONTINUED * itiiiiiti****ttititiit**tRk****tiiflittit*iu**tii***i*** 11 CONTINUE RETURN 12 CONTINUE RETURN 13 CONTINUE RETURN END 7 P : CARD fittiiii*****iii*iiti**i*******i******i**i**t*iit******i CCTTROL CARD TITLE CARD MUDT ARRAY CARD VCST ARRAY CARD CQHSTS ARRAY CARDS IR‘ ARRAY CARD ISMIN ARRAY CARD INITIAL ESTIMATE CARD DATA CARDS t**i***i*t*ti*fi**tk************i*t************i**tt**** BLANK CARD 6 ‘9 CARD 271 _ (M+.x') - ‘ “_—:—_— —2 1p (M+)(X H: (1B ) and Kt = —£§:*¥:*§:l (13-3) (M .X )(X‘) . in which Kt is the anionic triple—ion formation constant and (X_.M+.X_) is the molar concentration of the triple—ion. Other symbols have the same meanings as before. Equal activity coefficients are considered for X' and X_.M+.X_. The mean activity coefficient, vi, can be obtained from the DebyefHUCkel equation (Equation 1A53). The mass balance and charge balance equations are, + + _ — — 00*: (M ) + (M .x ) + (X .M+.X ) (X') + (M+.x‘) + 2(x‘.M+.x‘) (lB—4) (M+) = (x‘) + (x‘.M+.x‘) (1B-5) in which CO is the total concentration of the salt. Sub— stituting for (M+) and (X‘) from Equations lB—4 and 5 into the triple—ion formation constant expression gives, 272 _ (X.M.X) Kt ' [CO-(M)—(X.M.X)l[(M)-(X.M.x)7 (13’6) in which charges are omitted for simplicity. Equation lB-6 can be solved for (X.M.X) as a function of CO, Kt’ and (M), . (Ktco+l):\/(Ktco+l)2-4K§(M)[C -(M>J (x.M.X) = O (lB-7l 2Kt The negative root has to be chosen to satisfy the boundary conditions. Similarly, substituting for (M.X) and (X’) from Equations lB-4 and 5 into the ion-pair formation constant expression yields, Kip = Co-(M)-(x.M.x) 2 (lB-8) (M)[(M)-(X.M.X)lyi This equation can be solved for (M) as a function of Kip’ CO, and (X.M.X), 2 ‘ 2 2 2 Ki (x,M,x)Yi—1+\/[Kip(X.M.X)yi-l] +4KipyiECO-(X.1VI.X)] (M): J 2 2K Y ip i (lB-9) 273 The concentration of M+.X_ is given by, (MX) = CO—(M)—(X.M.X) (lB—lO) The concentrations of the species can be obtained by an iteration technique. Starting with initial estimates of zero and unity for (X.M.X) and Y the first value of (M) +3 was obtained from Equation (lB—9) and used to calculate (MX), and more accurate estimates of y: and (X.M.X) from Equations (lA—3) and (lBO7) respectively. This procedure was repeated until convergence occurred. Then the final values of (M), (M.X) and (X.M.X) were used to calculate the relative mole fractions, Xi“ The values of the ad- justable parameters were then obtained by fitting the calculated chemical shifts to the values obtained experi— mentally according to the following equation, = 2X16. ' (lB-ll) The ion—pair formation constant at each temperature, (Kip)T can be expressed as, o AHip 1 I 1 _ _ — _ lB—l2 (Kip)T ‘ (Kip)298.15 exp[ R (T 29 15)] < ) in which AHip is the enthalpy of the formation of the ion— pair. Four parameters were adjusted for each salt. These are, 274 CsI CsBPhu (Kip)298.15 = U ’ U(3) Ang = U(2) , ' Um 6M.X = U(5) , U(7) 6X.M.X. = U<6> ’ U<8> Triple—ion formation constants were calculated from the Fuoss equation (l-8) and used as constants. The chemical shift of the free cesium cation at various temperatures was chosen from the CsSCN data. The FORTRAN expression for this problem is listed on the next page. C. Ion—Pairs and Two Kinds of Triple Ions The equilibrium for the formation of ion—pairs and two kinds of triple ions can be written as K. M+ + x‘ ip M+.X‘ Kt +_ M .X- + X_ #: X-.M .X (lC—l) K t _ + M+.X‘ + M+ z M+.X .M where _ (M.X) (lC—2) ip (mom: t * FROM 275 0F TRIPLE-ION *iikiiiifittttit*t*ii****tt**tit**i*itiittitiii*****tiii KIND QUEyMULTIPLE DATA SET 3. l\ ‘MR TECHNI *t**i*fi*********tiiti********t**i*i**ii****t**t****itii ION'LATR AND ONE * * RD. 99 C C AYTT q C ) 5 C VTIDF \I C _ l C 011 90 9.0 C R u C N 9TL )1 C U 8 C 98 9 9 1‘ C T. T Q; C TPTT. 2V: 99 C A N 2 S C PESP 990)) C R A 81 C OoKU )C900 P C E T T E. C NR 9M 41;)55 v: C D. S A 18 S C ’AL 9 ((0(( T C M N CQ) N C RIAT XTlPY. T. C E C S N E2 50 C C9VA XC(YY C T C T 0 P TPI C NOGD 9..LYT.Y. C .1 N Tun ST .1. C .11qu )V 9X 9 ) C T N AG .0 ATAF C XSE9 09)R) 3 C N 0 TN EI F APL C 9E 9T 0‘10 90 l C A .15 ST. LT 0N .P C PROA 310.15 9 C T Tl NR PA ONPI C A 9TB 9230‘ 2 C S A» 0 01 IC SIOOR C LV9N 49(5T 1 C N MSTCA R RNTIITW C 9AM.L 9 (..UTI‘D: 9 C O rug... 5 TLIOAHT OTC T 97:... XESNC l C C DEAN. ALICQFFIFC kI9S T(XIL 1 C FzTON FTPT COOTIC 19PM XHP9M9 9 CEC E810 OE.CE AHC 9T 90 X 9 9)S\l nu CRT. ..HRNTIT M AFC—LQVQCRSC ..LSUT. vn))UInu 1... CUP OUOA wa NRRCTTT C PEF 9 9532 95 9 CTT ITCMFEOFOFFQFFNLC ATQT T20()( 9 CAC .A ROIIOI FIIEAC T 905 PlwaLCT 9 CRE EPWO CT rLF HHCCC IuXth CUI‘AEF cu CELCLFCFSIANF.L2CISSK.IC 9T. 9N J 9UV(0 9 CD E/PPI EFROOO O OHC 0 END: 9 H 9)FGTM 7 CMIlIMTRIFTI MT LLPELC 4 FOOT TT09IS9)9 CED*FEAIWENTN:FTAA HC T AXFC. EPU)EC\IIJ1 CTlxtTITM ALOEAOAIFCCTCC F TA9E W0309N019 C:: : RPACCRIGiIIIC C n N NIMIV IK9G)05(5 C)) )E0.H FTTESHMVADC I I G9Tl9 //430)(X9 C12 3CFNTYONAM SEESEC LoK ETIXY TT((26XX4 C99 9N ONTCEROL HP VC [0: N 900 DNXO(1RX 9 CTT TERIEI:CT 9ALCCLRC05 A64 EUU99SEI F09193 CAA ARlszNNhACAzzAECol ZLT NO 9PYTRON 9T0 9) 9 , COD DEA))ILUETTC))TSC1 a DA 9:! - TLKXY 9SFPO) 95),.02Enul CL C “UnU ECICJ JFrH34TACCEVr178OFTC+8 RH4N3 T 9TJN//IO)(01(U6628 U U o a UC(( (E.(I\CC:N«D_LM..((TCCSQJ AHKTI: o UNKXJONNSGCOSC‘Cl‘ N: __ __ :NNWN 10NNCSS CURNUHUATNO 9HEUU : :CTnd C9FKR 4 OONRUCOON5OV.TI\TOIEF._D.KRIR.1..U: :RICTT TI:O99: 9MCACH99)\ICT: RN 99 T RMU 999MVHE(ZQOQI¢.VTTD.PANUTUT:ALUTCCVS SF.1\I\)/ALE:H__C\I)ln/LC:F DECIEG 0F BMOXMYMVWMDri‘NlN NAAVUTNTNQCMATNCNN Nr._:lncM..A 9XPR:56(I\CSE DDRNKYGTN NUUONX QYOOIOLGCXOOOTTOOWLOEOTACEOCOO ORK(I\ACMVMLMM((XXCTR RRON9OLRI RSC9QLYCCDFZCDCGCIJNNRCRCIGTRCCCC CTFUUGTEEADDUUXYCIT AAWAthAK A 1234 12345 C C CC BEnLEDr 9 C 1 7 «C 2C AC 8 9U 9 92D 0 C C Cn SLTLTLD ......C c C C NCEAZLATAOC cw. C C PiPHRHFMLL? CCCCCCCCCCCCCCCCCCCCC l 'k flu...»— on," y. .. *it‘kfii‘k‘k’kiii*t‘kfi'ii‘kii‘fit‘kiii‘ktt ULTIPLE DATA SET-CONTINUED 276 UEQM tank r Lx MMR TECHNI ION-PAIR AND ONE KIND OF TRIPLE-ION FROM *ttfitttttifiifiiiitittit ii*****t******t*‘kii‘ki‘ki*******i*~ki**ii**i**~ki***‘k‘k****i ‘k ‘fi ‘1 ) \I ) L L AA A C ) C T ) T . E _ ) I.‘ ) l ) U l ) ( \I * I‘ \I \l \I X 2 \l A \I \l X \l n... ) ) X t 3 G ) ) X 1; * 7 L ( i t) E 7 L ( *)* 8 A ... A tT M 8 A ... *TA 9 C \I M. TD 0 9 C ) TDM o T ) A Us I\ u T \1 03..., 1 _ 2 G S/ + 1 _ 2 S/G ) / \l nvt I\ (\l \l / ) .rlufi ()l\ F ) 1 )0t * /) l ) ) l )o* /)* FL rK l\ \I ) TIA K \IN C r: R Ix \I ) TlAH \Ixh'n in. Cr VA .5 \l D: _ M D: )MI .1 ( D: X Cu ) D: _ M: \lAw. r T 01 M” X S 1 *)A 1 NE U5 M X S l *)A Mr: / 0C E ( * :\ 0N6 07M( 0 *3 E I. ... ( DING M1107 O 2: T *7C X 70L(705ET ) T A T *70 X 70L(7CT03 ) a .11. t )5 o X 32!n*32 (Du f. )TC .w )3 o X Tidahirorxnnnx. h. ) l ) S 2 . (CK (OTC M NO FET ) S 2 . (CK TQ(O M 1)( 00 2 SO( M O/TPO/TRS E LG 15 4 SO( M O/TPORS/T rr. QJL- TT. (‘1 *T/ E T)... 1T.) 51.... / A) \I hi..l\rU (\l *T/ E T.)* *Tfiu...) / T 9) U? U: ) 0(sl ‘1)0 )037 ) C1 1.. T+G U74 nun: \; U() ))\1. S7.)O ) ATP 00 _ 9) o .0) 0*3062 oODFGtZJ)M T0) ( *)\I . 9) o .0) oinUAUAJZ 00*3DAU)MH Dank—Cree (TALIQAUS 1)tG(*4GPu\anauF—L _nUl)x 4M1 (Tn/...anbs 1)*GI\* AGOCJP..H)FCL .dnUILfL)) (At—1.x)? +MU)T*I\\I(O 0 IF- NU‘1.X n2D_ (At .()Dn +~C)Til\) o o‘turr— («uTJH77 FCi)(n_( \lrL oORAI‘OT taro/N Mu oX(/ 2... a PD*)(U( )E ‘ORA(UC6T o/h. S(/T o - XdAL+ 0T. )t 4. oQM+ cannon/DEM. ECXXN oAE XIuAL+ :1 )* u. oQM... ca. DRUEM TSU . ET. C(MHAA)UR 1)(USA\I3Q .62EE _ o/XL .UHrWC EIxMA)nUR l)(FJSA)Cb2Am lLE ST cPLG (SACE 1U (2+ 0 . U2 .SE7((( )TN/«AZ _ u- 0L (SACZ .9 (2+ - _ 52 u7r\.,\:L(( NSlN . oEtTGTtESEXt)EAH(*E+L9(D.( lGMXCT‘LHfl EtTGT*ESEXx\IEAI\tEQ/(+er( ON‘ETTU)S(* tL+UXi2L298,15 expE— Rip (l = l )J (lC-lll) The subroutine EQN for use with the KINFIT program is given on the next page. It should be noted that all equations and iteration methods described in these Appendices were checked by testing the numerical values of all concentrations and activity coefficients for consistency with the equilibrium constants and conservation conditions. i i W 7 “ **** RP99 C C ... .w AVITT 9 C ) 5 C W, ... .. VTDP ) C E l 8 SC i ... .. 01.90 9)... C R c E NC 3 * * V9TL )1 C U 8 I CC 1 * t 9599 l( C T T 9 C IC g * a. TPTT 2Y19 C A N 2 E5 .C m i t PESP 99D)) C R A P1 LC *M ... OqXO )0 900 P C E T T S . PC w *0 .. ....R 9M 42.25: Y C P S I.» 8 ....C *AK ... 9A]. 9 ((0(( T C M N F01 RC ...Er ... RIAT XTIPY I C ..L 0 Tab 9.02 STC i .x C 9VA XC(VIVu C T C T .N TD. C ...N ... NDGD QEYTY C * TN OST ICC iOTt IIIJ )V9X9 ) C T N AAG INATAIC *IEt XSE9 nUI).K\I 3 C H... 0 TN. R _0 APNC i.S* 9E9T U)99U l C A IETSI I EIN .OC tr: ... DROA 34in )5 9 C T TlNNR AN LTOvNaNIC *LA* A9TD 9230‘ 2 C S AoAUI PO PCIOONC *PTi LV9N 49(5T 1 C N M8TCA .I IfiTIIAC 1.1/Ht 9AL9 (fiuTQP 9 ”C C ROS C. NT HKRAT C tRDi T 9ZT. XafiCuNmO 1 C C OOLWN. 0A TFCAFFC *T i “I 98 T‘VAIL 1 C F__CON 1C COOC ‘1 ... E... I 9PN XHwD: 9M. 9 9 CEC ECIO FEE C C! *FLt 9T90 X99)S) 9 CR1 1R TITFLNOLEESSC E #00:... COQUC X))~.UIC 1... CUDI. OJMAP CCAO OBKETTC R l t 1* PEF9 900295 9 CTT ITOMFE TINMFRFFC T 8 *ST* AT9T T20()( 9 CAC .AIRCINENO: FIIC I 2 iDLt T905 P(3LGT 9 CRE ERTO COMAIAF HHC 0 thumb... .UVAFC CL‘AHCVC. Q. CEL LL—LLCEICI TCCF:¢..C 0 *IM... 9T. 9N .U 9UV(O 9 CPPL/PPM EFTFFAE O C \i 1 *K9* . EUP9 H9)FGTM 7 CMIlIMRRIFAOORMT LLC 1 ( ... E* a. D. 90.] TT0 9.135 9) 9 CE U... REQIPEDu TCFTIIAAC 9. . ..CU... T. ..HXFC CPU)_LC\I51 CTlxiTTFAHLOTNKJh. oIFCPvC T. ) ... ”HG..— F TA 9E M030 9V01 9 C: : : PAnvNOOEAHIIIC A 0.. *T1* o N NIMIV IK93)95(5 C)) )EP.H EIICTShMMC D NF * N... T. I G 9T1 9 //4.2J.L\lr\x 9 C12 3CINTYCTTNE SEEC J FCC *SHt LoK ETIXY TT((26XX4 C99 9NA0NTNAAOTL HHC ( TR «NC* E0: N9DD DNXO(1RX9 CTT TEPIEIORRC9ALCCC $5/T .1 ..ME* AG“. . EUU 9 9nQEI F0 91 93 CAA AR. __ :VCTT: ACA»: :C T18— ... T... Z LT NO 9PYTRO.N 9T0 9) 9 COO DEN)\II:NNXTIC))C13 o .0. *R * DAsF. TKXY9SFPO)95)62E13 1 E E 0 ECJJ JF034TNEEMCM178C+N81N *IR... HRH/4N3 T 9TJN//IG)(01(U6621 U U UC(I\ (EI(1\CXCCXBrLMJ(I\CSOQ/(r_ .x «hum... AKTI: o UNKXIUAUNNSUOSRu‘ N: : : :NNNN INNCSS SR :UUAMNNE 9HEUUACTCACZT iDNi C9FKR I4. «UnUNnnIUCOOvV 3.JT(TOIEERKRIRI(U:RICTT TIHN 992500 9APCH99C7.::DH: *— 1 Phi? T D.MU999MNE(3SISTTP?ANUTUT:AUTCSS SFO)}AQCCMH:C))C:PFPF *N .x DEIEO or: BVHOVAVHV..M.M..V..P(N1N NAAVUTNTNSMTNCNN NELIlnéuxhX: :anR__quDCSVPLM-T. to * DDRNKYGTN DUAUNX 9YOOIOLOXOOOTTOOEOEOTAFLOCOO OR.I(I\AMMXMLMM(I\CTESULD ... I t R.,.KON 9DLDnvl RSC 9 9LYCCDFZCDCGCJIH..NDxCRCICDnCCCnL CTD.UUGFL_LE_LADDUUCITTTT .x * A«HH...AN.* :AK A 1:2/2&4. 19:34.3 C CV i * CC BRLBPQ C l 7 .8 2C C i * S9U99:U- C C .t * C:_._(.._LT.LN._.rAO r. P» C ... i .\C». AEATAOG Q. C C ... * ... t PJPHRHFMLLT CCCACCCCC~CCCCCCCCCCCC i * FROM CF TRIPLE-ION NMR TECHNIQUEoMULTIFLE DATA SET-CONTINUED *iiii‘it*******tti'kti‘k*titfittiii'iiiifiitfi*iiit**titt***ii *iiitiiitiitttttiittt*ttiitttiittiiii*ttittttttittitttt ION-PAIR AND THO KINDS ‘I i ) 7 8 \l 9 0 O c 1 1 I. + \l ) R M P P: M... t ..L N .1) T. 0 ll 1.. ... I ‘(00 \l TI XXUO 2 P .1 XXnul ( ... .3 * *1A “L n.» 51 0 . o )0 0007.! ( 3 MT 0 CT ()I\ A 00 Pl“ 6»... .. :00 X1/ *3 HMGG E‘) N) CF?!) ...UM O... .))7%I abir. 1;.) 1.1 o o )NN . 2 106 o 2LT. 10.U)* P QR 3......LG. (III... (EA ENOOCUII‘FE/L‘ o oTTU..D.P.XM..UX 0T ) M X v. D ... A G E M \l O 2 ( 9 + T- \a A x \l \l D M 7 O J x 8 \I O l\ D 9 0 1 Q. ... o o + T. A» 1 1 \I S) T I 4 NH NB E \l \l E on T R M * C9 ( P E N ,0 + M... t 0 \a a nu ‘1 F. N I ))1 .0 Y. T O T 7N..\.l\ \l 1 MAD * I 0. SEE... N at. ) T. .1 RR, X 0 ... a. Dr 7. a OTTvJ MM .1. LC 1.. i 7... o T89... E \I ) TT. .U Q. 3 ((fi / 0 F F r... . o )0 (M OTT) ) G I I 80 l\ 3 Mil. (E GRR2 x \I D D (G ()|\ A It \Ihuauu 9 Mr 0 \l TI TI .7 Dull! 60 )N 088T.) pr. 1 )1 t t )\I X1, *G "X 0* *AC - .\l\l1..l\ Q. \I M1 E‘) N) EM“ MIOCSDA N E11¢X 2 6 D. *Uum CC ..L Xooclu/ X 0((Xx 2 ( ... 0 Rip... 1.) )* MECS‘AAQ." IXXX/ 0 U Ar. )NMm . 2 71.06 1N EL“ iSAiE oXX/M C 1| HIHC 300).! P QR (0M+069T(*( T/lxxs.)+ )POL (111* (EA XIXMN?2SPA(NGMXNM.7(:.RX7LHA EUII‘AE/L‘ XTMEEQI. oNXumSX oEMXE «+(quCATC U_.D.D.XMUX 9T .(0M 9 orDG/TI.‘ t ( ) N))/CONSTS(JDAT92) 100+§9) 2C0 DIG. . 3 RE RE )* 01:15.... Tc..c..... (C... OTT) GRRZ )Qru 9 OSST) 9.. ... AC MUGBCL X o . oJ/ ).. M.Env5(nuncur .lN rfL4*Q7A*E MX)/£FXN) ’ -9- E.)/((3.0*PTION*EM)+1o0) 7 f N ( - u E-XN*EM XN ToluE’lG’GO T0 XIXM. N7ULSPAHI\NnC XTMHEtzw/ ONXVISX o ADSAAAKR: ..XANMGR ‘PE:R100EABMT:ErL._rURUM.nUU(E:nN-NRZ..XANMGR (PE..R100EAEV-T TTUUIMNNMH‘G~1£RGMH .. Z ..nNTaR..C:GAHEAA 2 ...HIDM:nU: .. ..N)RIHHNN(GIER..UM: : :NTQQ.C:GAFCA IIIUIUTOOO: ..T. I ACQEMmNXF.S . ((AH _. ._ ..RHAAPO_.DR: M.XC12UTAUOO: :T. ..ASENMXES . ((A: .. : F (((‘NIIIXMNG( : : XXMR(: : : MMHTX(PTT...LR:XXXML((TNIIIXMNG( : : XXMR( : : .. MM~TX( FFFFOIITM.ACRFMXMMXTFABCAAAMFLEEMMu.MVNXAFXEOIITM.AORFMXMMXTFABCAAAVF IIIICPPPEGCAIEEEEESIAAAGGREIABTODDDODDCIXRCPPPEGCAIEEEEESIAAAGGR...I .U .C 0 n.. ...... an. 1 1 0 0 0 .. 2 l liflfi 283 i * it*fi*i*t**i*tifittii***t*****i******i + BETA*DMX)+(TETA*DXMX)+(OVEGA*DMXN) 0 TO 35 THO KINDS OF TRIPLE-ION FROM MR TECHNIQUEvMULTIFLE DATA SET-CONTINUED ititifiiii**fiii***i******i****ii****ii**itiititittfiiittfi \I \I F F I I D C ‘G \l ..I T. )1; t t )) \l)1( 4. ‘1 M1 *11l\x 2 O, D- C *(‘XX 2 ( to N tXXX/ 0 ..U AC A *XX/M nu ( \IHNC *I/XXS.)+ UPoL R *MWXM-M7(8RX1LHA I ...—LVMXE 0+(XM(ATC A * Dust... _ i N t O t I ... .. i * ... * ...tt MQEX9EMX9EXMX9EMXM96AMAQJCAT 1. F. 0 95 HH~)I 61.. 9 (x 0X5 TX 9 ) )a- Zone o (600 X 31 VA) 9E .2:L7 C 0P 9 ‘Erflx O). MID 0. 0G EAINTSEptrL U 5 3 C e.u(U 0C . 039 36 M9EX9EMX9EXMX9EMXM9GAMA9UDAT X 90 )4d nu 0( COT 41.0. 9CD 5E7” €0.90 nUAHVrrU EoTSE U1J‘ o 7 3 o U 4 23 ONEO-l) GO TO 345 ..L U U U U U U ..EE..nURUMhfiUUI‘E:NN:P(TNSSNN=(TSNNNNNNENNNNNNNNNNNKN A 2 2 Arbum :D 2 ._ ZW...)DuIDAEAITITRIDPLATRIRIDH.IMRIRIRIihIRIRIP «U 9 AU 1 0L 5 2 CL C r. 1 1; FL E HAAG 2 DP. ._ MXC12UTILTIHTIIUTITIMIUTUTUTIUTUTUTUTUTUTU PTTER .. xXXML((TNSI‘IRNI\I\TNSIR(TNTNTN‘TNTNTNTNTNTNTD LE.LMMwM MMMXAF,XEOPLFROOFFEOEROFF.O—LO.LOFEOEOPLOEOEOCOEMN. AET:0.DDnUDDDCIXnK.PwRIUFCIIRCRdFIRCRPuRCIRCDucpcafirbRCRALRE 1; 1 7 5‘. iiiitiiitt**********ii****it****i******************i bNK CARD S D R DE DOA RT ...FRCDAAH AAanCM“ CCYA I ACVIT DYVIR ASS RAARVIRED ARRAAR R CRR RALA AHASR AC E TAN]. LTTS ITA - ...-.. $5.)» .17. a “TROL CARD 0 CARD tittitt*tiittttttifittit*t*****it****t***t*t*tiiittitttt CC IOCCFSKiaL 7 THNCIIID * Be 9 9 CARD APPENDIX 2 DETERMINATION OF ION—PAIR FORMATION CONSTANTS BY CONDUCTANCE MEASUREMENTS; DESCRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN A.' The Onsager Limiting Law The Onsager limiting law for lzl weak electrolytes is expressed as A = d(Ao — s/Ea) (2A-1) The ion—pair formation constant can be written as — 1'“ 2A—l KA ‘ __§_§ ( ) Cd Vi which gives I 2 — l+AK C a = _E:_____A_:£ <2A-3> 2 2KACY: in which a is the degree of dissociation of the ion—pair and, Y+ = exp _ H.2Ol79XlOO/EE \ (2A_u) 50293 /E 3/2 (DT) [1+ (DT)1/2 281.l 285 In these equations A and A0 are the measured and infinite dilution equivalent conductances respectively, 3 is the distance of closest approach in Angstroms, and S = 0””‘o + 5* (2A-5) where 2 a* = _.___§_£_____ 6DkT Vim-429m >33” r". 'U 021 Mb 0 5 In HC VA 0 xkflfir c 1‘). {'71. vx Hmf'nx ")NOA o N N HCHT<-47H~4CNHC3 + (Amml — so. ( m m U! ONE-’1) ac~4c4 :n I”: R ‘0 '4’: C "1m H (T) p.) CCCCC-{CCCCALOOC H rd (N n) H m P1 m mHTUHXJHmHmH70H‘DXt-4D—4,V,‘HIU—-4DMDUHIJVTF. U; C ("NC—IXNWDH 'annflninmnwnmnflnmfu U") xix—«nu? zmomomomomomom‘nomomomoop11momx11>n oqzqzazazqzazaazazazqzmHAmZAAAPD cchcqca 2222224222322m22232224avHz?nmu Q CARD .OL CIHU E CARD TfiJT CARD IQL ESTIVATE CPRD A CARDS «amt-*4 H T P I "T i L NK CARD M 9CAHD l)-S*(SGRT(APG))) GO TO 33 ) T0 45 )CANDoSTPEng ALPHA a) 60 TO 2’ *ittfiit‘kittti’ifiti****itt*t*****i**fi******ii***i*it*i*t *******************ttiititifiifiitttfiti‘hiiitttittit‘tit 289 in WhiCh E, J1, and J2 are the coefficients whic have dif— ferent values according to different theories. Other sym— bols have the same meanings as before. This equation can be rearranged for use with the KINFIT program as, AO—S(Cd)l/2+ECleg(Cd)+Jl(Cd)—J2(Cd)3/2 l+KA(Ca)Yi where -l+ Vl+uKACyE a = ————2——‘ (213-3) 2KACYi and u 20179x106/EE u = exp ' o/__ (213—4) .2 (DT)3/2[l+50 Tii/ga] The ion—pair formation constant and A0 can be obtained by an interation technique similar to that which was described in Section A. The values of the coefficients according to different theories are, 290 (i) Pitt's Equation Linearized bx_Fernandez-Pirini E = Ele + E2 E _ K2a2b2 l - 25C E _ KabB* 2 1601/2 and J1 = Ole + 02 2 _ (bKa) 2 ol - 12C [in n(——7§) + — + 1.7718] * * o — 3—53 + B bKa [0.01387 — 2n 2 ' 1/2 ‘17? “16%” C 8C C _ (bKaL3 [1.2929 + 1.5732] 03 — 6C3;2 b2 b ~w‘ 2 2 _ 8*(Ka) B*b(Ka) 291 where In all expressions a has units of gm. Other symbols have their usual meanings. (i3 Fuoss-Hsia Equation Linearized by Fernéndez-Pirini E = The same as Pitt's equation. J _ (Kab)2 2 2 01 - —:flfiT—[l. 81U7 + 2Rn(* +—§(2b +2b-l)] ) C1/2 b 02=d*8*+8*( Kab 01/2)- 8*1601/2[l° 5337+; —+ 2n(——7—)J 2 3 b (Ka) u.u7u8 3.828u o = ———————[O.609U+ + J 3 21:03/2 b b2 * b on = 8 2i: )2 [— %(2b2+2b 1) - 1. 938”] + 2 2 8*(Ka) B*b(Ka) “*B*(ET72) + 0 ‘ 160 2 (1.5405 + ELEZQl) 8* Kab [1* -2.2lgu] b - 1611001/2 §5 292 i*******t*i*i***fi******* i CUCTANCE ATIMEFORCON *iiit*iiiitiiitti*tiii**********ii****i plTTqS ECU fitti‘k‘k‘kiflk‘ki*‘k‘k‘k‘k‘kiiti‘k‘k‘kii‘k‘k‘kii *‘****i********* * DEIEO arr DDRNKYGTN RDAON 901:.th rArMHWAN‘t :AK CC BRLBP9 S 9U 9 920C - NCAfiEATAOG PdPHRHFMLL? Us . D R A C rt. N RD. 9 9 C O C AYTT 9 C N I C VTDD ) C 2 O T C 0190 90 C / I A C N QTL )1 C 1 TI U C 93 9 9 1.( C ... U 0 C TPTT 2V: 9 9 C t L F. C FESP 9.0)) C ) I C 0 9X0 )0 '03 D: C E D E C NR 9M 49.)»??? V- C P. C C 9AL Q. ((nUl\l\ T C U E WN nC DnleTI XTIPY I C T T. AA C CQVA XC(YY C A M. I LT EC NDGD 9EYTY C R C N C CC IIIIU )MV 9.x 9 \l C E I GU NC VASE! 09)R) 5. C P N. F ND AC 9E 9T. 0’0 9C 1 CTN.” I N IN TC pROb 3410):; Q. CNF. TI TIO «LC A 9TB 9230‘ 2 CE?! RN 1C UC LV 9N 4. 9(P3Tn 1 CI... EAT: V DC 9AL9 (3T1? 9 CCT TTA IN NC T 97:]. XGLQVNO l CIPM ES L1 or INT. 9C9 T(XIL l CFAHCLVVNE—t CC I 9PMH XHD. 9M 9 9 CFTSAOCCFS C 97.. 90 VA 9 9’8) 0 CPLSIRCMNNCT TC ESUC X))DIC 1 CONOA AIM ...uNNC PEFQ 903295 9 CCOPPNTTSEOEC T AT. 9T TOLfiLl‘)( 9 C C OCSTIILC C T 905 D.(7JLUT 9 CV. NEIUIH‘HCTAC S 2 ) JXFC 0W(A5P P CTCICTCCEILVC / T 3% 9T QN J9UV(O 9 C11 KAN IFRIC \I D TY END. 9 H9)FGTM 7 CVPYAMOECFTUC x / Dr. Dr 907! TTnu 91$ 9) 9 CITTTRCSIENDJC E \I / t AXFC EPC)EC)51 CTCISO RFOEEC t 3 )T TA. 9E “0730 QNOI 9 “VCESIFTEE CC .u \I . 53 .NIMIV 1K onu) 95(5 CALOD NVEINDC ) 8 E 7.5 ~.\ 9711.. 9 l/41v0)l.x 9 CZCLC RFNnUn/ZUEC ‘1 EA7. r31\ ETIXY T1((2€XX4 CHISNILICJCRC) )1 GIAO A/ N QDD CNY/Ol\lp\x 9. CCLCIOAAH :E UCl 2(\12*2 nUb EUUQOSEI F09I93 CN(VIPVFA0TSC(23(X3¢AC 28 NO 9PYTROAV 9Tb... 9) 9 CA: I : ~IYT1LACT* *TXl‘fiunrl on». IKXY0SFDO)95)62E1Q E E3 ECM)))NUBEJAECS**S(T5K7L8g T 9TJN//Inu)l\nul(U/br022 .. UCA12SORQFZUEOJMnuNTTu....TIS(_L,DA(2 UNKXIVONNSOOSS‘ N..:::NNNN1NNCG(((IED 9 9: ..COUDORN(: 0/28 OONRJCOON3CT¢TOIEERKPIRI:RICoTTTzz:A2))CCSSCGC:AIEA: EMUocvMME(3$ISTTPPANUTUTAUTCASSS))AHE12C::::SCAA(AHA BMOXMYMMMP(N1N NAAVUTNTNMTNCMNNNIZPPo((CT23A::PP::PT UONX 9YOOIOLOXOOOTTGOFLOECAPLOCAOOO((KL1XXCDTTTXAKKBD_L.L 9Cc9LYCCDFZCDCGCJINN?CRCGRCCGCCCUUEAEXXCSDDEEAEEABAE 123a 12345 C C 1 7 8 2C C C C C C C AC CCCCCCPCCCCCCC 1' *‘k'k'ki *‘k‘l‘k‘ktt‘k‘k‘kfii -CONTTNUED t******fitt*** **3)))*((1- NCE *** 1))+((0-23484*BETA* ‘TA*AB*EKPA)/(8.0*EX))* ********tfi*‘kit* NDUCTA i***** /(12¢U*XX(I)))*(ALOG(EKPAA/ /(6.0*(EX 293 *‘k'ki'i‘k‘ki-kirt UATTON FOR CO ************* PITT98 E8 ***********i***** **i**‘k***‘k**‘k‘kt ‘h )M NA. R... ) TN M Sr... A (Dr. G T t 3 S \l T 6* 1 D O) I\ ,1) L1 X )(U) A‘ VA )Gw *U i T/V.“ N+ ‘1 nU)N 1 F.2d» )1 SRAH 0 R O M( ./TM .1 T1 AU )SA 31‘ 6* Rtfi. C */5 *0 TIEQ/ T 7.7.13 ) o S o) F.) 12 *lnh 0 ()0 (1‘ porn C #371 X/ A7A )t X) *16 \1 Rio i) 80. 0 THE )F Enid. 1... ST 1F ’94.}. — *S) ((DN2.” E 8(1 UTHUE 0(AM 0 8* . *RtRH7IMW o (2. ZUQ)T5PAE12_JE *oSlS(XGNo*)ENC 1'! GO TO 2 ‘2Equl,Ed2yDDQSTRENsGAMA9GAM Eo‘l) CARD 294 **fi**********i****************************t***********t * PITTeS EQUATION FOR CONDUCTANCE-CONTINUFD * **i***********iiii********************i**************** *********t***it***i***t**i**i*itittiit****i******t***** CONTROL CARD TI LE CARD CONSTANT CARS . INITIAL EST IMATE CARD DATA CARDc it***************t******iit**********ti***itfi********** BLANK CARD 6 295 * FOR CONDUCTANCE EQUéTIOV FUGSS-fiSIA *iitittttiitttttiitttittiittiittittittit*ittitttiitiiii *iiitii‘tiikiiiiitiiittiff*iiiitii***t*i**t*iiff‘kiiiiiit * nun... 9F 0 O‘HQNS QKTI: o CoFKQ 4 QM. 9 9 T PJPLIEAJ 0F. CDRNK.YGT.N RRCNqoLRI AAHaAN...” :AHK CC BURT—BO: 9 cu 9U 9 9...”... o r. r. (CLTLVJ. ..." r». .., CLARK. LTAHCC u. Plu.P.HD.\H.,FMH.LL7 RD. 9 9 .AVITT 9 VTHUP 9! CI 90 90 N 9TL )1 9Q; 9 9 ll. TPTT 2Y99 Burtr‘up 990)) fig. 9x0 )0 9n.n.\ Dr NR 9M “2):”:4 Y. 9AL 9 ((O(( T RIAHT XTIPY. I C 9V.“ VAC‘YY NnUGnC 9f._Y.TVu 1.1.1.0 )V 9X 9 \I VASE 9 0 9)R) 3 9E 9T 0)nu 9P... 1.. PRO...» 31991.3 9 A 9T0 9230‘ 2 Lv 9N a 9‘5T 1 9AL 9 (UT-(P 9 T 97..T XZOQMAO 1 ”I 9S T‘XIL 1 I 9PM“ xuwP 9M 9 9 9T. 90 X 9 9.13) 0 ESUC X)’3I.U 1 P??? 9 9002 9.5 9 AT. 9T T9..fi...l\)( c) T 908 pl‘SLALT 9 .1»ch fi..uu(.ur.p 8 9T 9N ..u 9UV(0 9 CHIP. 9 H 9)FGTM 7. P 9OT TTnu 9.13 9) 9 AXFC ED C)CC):,.1 TA 9F. MOQCO 9Nnul 9 NIMIV IK 90) 9c...(5 897.19 //a.&C)(X9 ETIXY TTI\(26XX4 N 90C DNXO‘IRX 9 EUU 9 955.... F3 9T. 93 .NO 9PYTCHON. 9T3“ 9) O IKXV. 9$FD Pv) 95) bertlg EMOXMYMMM Pl‘MNIuH I...“ C O C C N I C C 9. 0 T C C / I A C C I T U C C ... U G C C t L E. C C ) I C C r... .b .rL C C R C C C U E UN C C T T AA C C .n as. I LT. CC C R C N C CC C E I GU NC C P N F NU AC CTN I N IN TC CHE TI T0 CC CET RN IC UC CI... EAT .M. DC CCT TTA IN «.....C CTN ES LI 0C CFAEMNEE CC CFTSAOCCFS C CESIRCNNOT TC CCNOA AA NNNC CCOPPNTTSEOEC C C OCSTIILC CY NEIUINCTAC CTCICTFACFIAVC CII NAN. IFRIC CVRYAMHOECFTUC CITTTPCSIENGC CTCISO RFOEEC CCESIFT....=.CC C CALOD NVE:NDC ) rL C:_.LC RENCECEC ) 27 COISHILICIUCRC \l )1. XiPJ CCDIOAA :E UC21232()Q..A—L*2 CN‘VIPVEA QTCuC* (* * (quzA/CO CA: : : .IYTILAC...T* *TX( airmail Ac Tu ml. 03 I ) I.» . E8)*EX)/FDT {LCM ))\INUBEIUAHECAHSTTS(T.AUPC V 7 T 9TJNI/Ird)(01(U662n/. UCAIOL300.LBESMCMNJDDNTSSKKF.6 UNKXJONNSCGSE‘ N::::NNNNINNCG(((IED99::CAOSSCRN(EE:o 00M?JCOCV50T(TOIEERKRIRI:PIC9TTT:::A2))CGC::CGO(::2I RVU99.MV§(3SISTTPPANUTUTAUTCQSSS))AHEIZC::23:SC:ACC( NAAVUTMTNMHTNCMN4NM412PD: 9((CMTTTA: 29.9 0.9 : DUONX 9YOOIAULOVAOGOTTOOF.OEODHFOCAOOFJ(I\KLIXXCA—UDWLTXAKKKKB . .SC 9 9LYCCH.FZCDCGCJTNNRCRCSPC «A. «In/:34. In....2.4.3 Pb L , CGCCCUUEAEXXCGSSSF.EAEE:.:..A C C 2pc C C C C C C C CC CC PVCCC CCCCC * COKDUCTANCE-CORTINUED 296 EQUATIOK F3“ *i******i***i*t****tiiifiiifittitititi**ii**titt*tittt*ti i*‘ktiifiiitttfiititiitt*tttitttitttititttttiittti‘ittttttt FUOSS-HSIA * I \l i / 1 ) ... \I l 8 6 a. 1 \I D 4 7. c. ...U 3 O 7 n4 1 PL) * 4 4 ) o 2 () .- ( o 4 2 o ...nu E + 4 8 l‘ 0.. ‘1.“ EU 7 4‘ 7C + o )6 o I\ 5 + 9. _.5 \I Nu... I ~J a. O nJ \I \I EN 0 Cu O, 1 4 5 ML RE 0 O 0 n 5 P... A TR 2 1 (D ‘1) o * Pu Tu ST 1| l\ O Bad-I. 0 * TI (8 l‘ * 0 BC‘ 0 \I D 6* + 2 ( 89* 3 1 S 0) \I E i i“) l\ 1| .9. L1 ‘1 l\ \l )Ez I x )6) A‘ 0 C - nU )iC 0 x )6“! i U o P \I 0 73A; 0 * Tl/E N‘ 1 K C) a *TV. .44 \l 0)“ 1E... . E D.) 2 *EE (4 )1 SPA «URU ) ( K) / EP( (C M‘ [TM 1T. 8 G E) ) B(* *1 AU ’85 91 C... rU *C \l I\ +\l )7... Gt Rte 0*(—: i «U II. AP 2v [‘.an ’.9 *0 T5,, TIE/3 n... .0 A ...IK i nUC 0 21v ) o S a) F.) .2...) 4 o .- EE 1 ) .Prc (a. 12 *1A 0(30 2712x4446 aw. ...Ll‘nLS )_,/..K.1. “SQ (ls Ag»..- Pa... )T (QTT/l o (GGP)(E/ ’16 XI A7A )3 +» O 1* fiébrznr* X” TD-IO'SU 41> Hill C" :0 '4!) ATE CAR] *i'k*i’*‘kiii’iiiii‘ki‘ki’it‘kiiii*****************i* r" 1} T’T‘IQH’D * ‘.":v #- "QHETH? “ Zibflmrd* -UtCJF'T 110* 5.3 :U‘ '10??? II- $$tUHOdO$ XI- ... s 298 (iii) FUOSS-Hgia_ Equation Corrected by Chen J1 and J2 = the same as the Fuoss—Haisa equation. The term EC lnC is replaced by 2 ECan = 12 O £n(Ka) - B u £n(Ka) to satisfy the Onsager reciprocal law. (iv) Justice Equation The equation is exactly the same as the Fuoss-Haisa equation except that the distance parameter is taken as the Bjerrum distance, The subroutines (EQN) of the conductance equations for use with the KINFIT are given on the following pages. t i FOP COMDUCTANCE CORRECTED 299 (OR JUSTICE METHOD ) ECUATICE hY CHEN *iiiiititi****iitiiit*ttttttttitttittiiit*tiitt*t*itttt FOUSS-HSIA tiiitit*********i***************i***i*tii‘kiiiitit‘kttiii 'k * AKTI: o C QFKR a. RN00 T DEIEOoF DDQNKYGTH RRON ODLPI AAHWQK.‘ :«kru. FCC ERLBP 9 SOU’QtDo pt, at?” LT. LN 01..» fl... ...... C .p A E e.u.—l A» 6...»... PJPHRHFWLL7 .hr RPQ! C C C AYTT Q C N I C VTDP ) C 2 0 T C OT. 00 90 C I I A C N ...-VI. )1 C 1 T U C OS I Q. 1‘ C .1 U G C TPTT 2Y9! C t L E C PESP 090)) C ) I C vao )0900 P C E D E C NRQM 42)55 Y C 9 C C CAL, ((0“ T C U E ”N C RIAT XTIPY I C T T AA C CQVA XC(YY C A M I LT EC NDGC 'EYTY C R C N C CC IIIJ )VQX’ ) C E I GU NC XSE’ 0 O’R) 3 C D- N F N0 AC QEIT 0’090 1 CT” I N IN TC PPOA 310’s 0 CHE T1 T0 CC AQTD 9233‘ 2 CET RN IC UC LVQN 49(5T 1 CI* EAT M DC OALO (OT‘P 9 CCT TTA IN NC quT X2SNC 1 CIN ES LI 0C “IDS T‘XIL 1 CFAEMNEE CC IOPN XNPQH’ 9 CFTSAOCCFS C QTOO X99)S) 0 CESIRCNNOT TC ESUC VA)\Inv.InU 1 CHVNOA AHA NNMHC PEFQ 000295 9 CCOPPNTTSEOEC ATQT T20()( 9 C C OCSTIILC T 2 T908 P‘SLUT 9 CY NEIUINCTAC D T dYFC CUCASP F. CTCICTSCEIAVC S D QTON d9UV‘0 9 C11 NAN IFRIC I 5 CUP! HQ)FGTM 7 CVRYAMOECFTUC ) / PQOT TTOQISQ}Q CITTTRCSIENQC X ) AXFC EP3)EC)51 CTCISO RFCEEC E 3 TAQE M0309N01! CCESIFTEFCC C t . NIMIV IK90)95(5 CALOD NVE2NDC ) E GoTlo //43O)(X9 CzEC RERCZOEC 8 7 ETIXY TT“26XX4 CUISNILICJCRC ’ )1 E Xig NVCD DNXO(1RXQ CEDIOAA :E UC21232()9LE*2 EUUOQSEI F09193 CN‘VIPVEA9T5Ct(**(X320/C0 NOQPYTRON9T39)9 CA:::.IYT1LAC*T**TX(othpl IKXY!SFPO)95)62£10 E EC ECM)))NUBEJAECASTTS(T8PPK7 T oTJN//10)(01(U6622 U U. UCA123OQEBESMCMNDONTSEKK56 UNKXJONNSCCSS( N::::NNNN1NNCG(((IED99::CA0850RN‘EE:o CCNQJCOON3CT(TOIEERKRIRI:RICoTTT:::A2))CGC::CQ0(::21 RMU99OMMF(3SISTTPPANUTUTAUTCASSS))AHEIZC::23:SC:ACC( fifloanMNNPCNlN NAAVUTNTNMTNCMNNNIZPP9((CMTTTA::PDPP: DUONX9Y0OIOLOXOOOTTOOEOECAECCAOOO((KLIXXCAUDDTXAKKKKB PSC9QLYCCDFZCDCGCJINNRCRCGQCCGCCCUUEAEXXCSSSSEEAEEEEA C 1234 12345 C 1 7 8 2C C C C C C C C CCCCCCCCCCCCC 300 * i 80 )-CCNTINUED EQUATION FOR CONDUCTANCE CCRFECTED (09 JUSTICE METH' FY CHEN FOUss-HSIA **u*********tii*tfi*********i********i*i******ttttt***** *i‘iti't‘kii'iiit‘ktiti‘kiiittiiiii****i******tiii’i’itittttiti i 'k ) I \l \I i l 1 \I \l ) I 8 6 4 3 _ ) C 4 7 9 t \l 3 O 7 n6 1 i 0 t 4 4 ) o 2 R O t l! o .4. 2 0 TI 2 B + 4 8 ( 2 S 1 ...C 7 l\ 3 .7 . l\ l ( 3 + 9 5 ) t ) I 3 4 o 0 ) ) 2 \l 0 5 9 1 4. B M. [U 2 o o 0 . 5 B A E ( 2 1 6 ))o i Q 3 ( U ls l\ O 921 0 t T .- t)( * 0.8C( o ) D ) )0+ 2 ( BP.‘ 3 1 Q» N A.) E t *K) ( ( i, E Pa.) l\ ) )EZ I x )6) R 0 K/C . 0 )*C 0 x )6” T o E); \r o Infinp o t TIE S 1 ()K C) 4 tTK “a ) D)N 1* . GAE DI) 2 OEE ‘4 )1 SRA 01 ) 0P( K) I 88( (G "( ITM 1d may LKPJ E) \l Bl\* *1 AU )SA E) R. AEnC *C \I (+) )2 Gi R*G O()Eu a.-. i .U *I:L AD..J /\8u )+ .0 715/ 7l+Mlu 2 «U DB’s-uh TK * 0c 0 23 \I o S 0) CA 03) 4020* EC *’0P6 (4 12 *1A OEGO O 2TAA?E*LC E(r,¢h)2vr1 U54 (( A0,”. 5+ ...T. T. (DTT/liAo (GGP)(E/ IIG X/ A7A )N +SEE)/Bi2 +01K2‘t) ’21 X) *16 ’REO 0 )/*8))BB( )LZE*(AB 2+2 *) 83. OTRG G ))T+2A(B+ AAO‘ttTE E2+ )F E2“ IST ) 253)*Ttt7 T*)*B)Et *4) IF )945 .t5) 2 ) *FS)*E)A4 E0))BIBA AG) ((DNZ—N E811 ( 1 *:,.I\n/_Q...EALD.1 D. 022(E*Tn T12 UTDFH I‘A OS). x . 9.4/(Dirt *K8 ind‘t/iAE EZ‘ *RiRnU‘MuHH 0‘1 0 VA 0 BOEU‘BtE. A‘thAHB B+U 0G)T5PAE21.(E . E (28*t8At1 H+*58TP( (1t oSlS‘XGN*0)UNC C N *04A2tPX( P)152EL( (43 4+(((E(A*T2(0L L o AGQOHCCKgi L)G(BPA( ABII‘ Ol‘l‘. (GGECCXT+:(MAG(+HA EA E E EH E E E E E E A.(2PPPE*1 .IIU+.XR0USAM¢U0TC UC U U UT U U U U U U /2:8LKK(AE (82(3::: :22N01:QoEBGAT(0E:NNzNNNNNNENNNNNNNNNNNNN B(A:AEE(T:):*(:(123)4:(Io.NSINA:GA:1M)RIDRIRIRIMRIRIRIRIRIRIR A:HA((((E1)2U:3+&44)44:Tl‘F:(A:AzPCCI2UTIUTUTUTIUTUTUTUTUTUTU :BPT::::BGBGolG)GGG)GG2N::RR:"TMM(L/((TNSTNTNTN(TNTNTNTNTNTNTD BBLESI2C(IEI$JIBIIIBIIJOFDTTGAAAAFA FXLOEEOEOEOFEOEOEOECEOEOEN Run/A95EEE(7...5Z(F_ZBZZZBZZECF DSSGGRGGIC IXRCRRCRCRCIRCRCRCRCRCRCRE l 1 1 1 1 1 1 5 3 4 5 5 0 3 2 1 9 0 1 2 3 1 1 1 1 CADD --.=.-. .‘._.-~. -.‘.--«v'. 301 *ii*****t*******t*t***********i***********************i * FOUS§-HSIA EQUATION FOP CONDUCTANCE CORRECTED * * :Y CHEN (OR JUSTICE METHOD )-CONTINUED * ********i**t******tit*********************i**********it **i*************t*****iti*****i*i*****iitit*****i*t**** COETROL CARD TITLE CARD CONSTANT CARD IfiITIAL ESTIMATE CARD OflTA CARDS ******it*****it****ii***t********i**************i****** ELfiNK CARD ' 6 . E 9 CARD APPENDIX 3 DETERMINATION OF COMPLEX FORMATION CONSTANTS BY THE NMR TECHNIQUE; DESCRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN A. lzl Complex Formation in Media of Low Dielectric Constant The equilibria involved in a solution which contains equimolar concentrations of a salt, MX, and a ligand, C, in a mdeium of low dielectric constant can be written as, M+.X‘.M+ ++ Kt M+ K + M+ + x‘ 4p M+.X' + X K+ X'.M+.X1 + + (3A-l) c 0 KC M '++ KX MC+ + x' E? MC+ X— in which MC+ and MC+.X_ are the 1:1 complex and the ion- paired complex respectively. The other symbols have the 302 303 same meanings as in Appendix 1. The equilibrium constants for these reactions are (M.x) ip = vi (3A-2) Kt = (iiefi i; = égigiflg> (3A‘3) Kx = z§¥§S§c> (3A'u) ‘KA = (M(CM)C(.XX))Yi (3 A_ 5) Kc = (%%%%7 = Kiifx (3A-6) in which the activity coefficients of M, X.M.x; M.X.M are considered equal and the formation constants of the two kinds of triple—ions are taken to be the same. The mean activity coefficient is expressed as 6 Y+ = eXp (-u.193:AxlO 5g<:;o+/:xwx: )) (3A_7) ’ 3 . a X + XMX (UT) [1 + 1/2 (DT) 30A The mass balance and charge balance equations are, 3 II C) II (M) + (M.X) + (X.M.X) + 2(M.X.M) + (MC) + (MC.X) t o (3A-8) Xt = C0 = (X) + (M.X) + 2(X.M.X)-+(M.X.M) + (MC.X) (BA-9) Ct = C0 = (C) + (MC) + (MC.X) (3A-lO) (M) + (MC) + (M.X.M) = (X) + (X.M.X) (3A-ll) in which M X and Ct are the total concentrations of the t’ t’ cation, anion and the ligand, respectively. The solution to the above equations can be obtained by an iteration method based on the major species and successive corrections for minor species in solution. If we assume that (M), (X.M.X), and (M.X.M) are small, the major equilibria are, Kx M.X + C 2 MC.X (3A—12) and KA MC + X it MC.X (3A-13) The mass balance equations can be written as, (MX) + (MC) + (MC.X) = cs z CO (3A-1A) (C) + (MC) + (MC.X) = C0 (3A-15) (X) + (M.x) + (MC.X) = cg 2 CO (3A-l6) 305 This set of equations yields (MK) 2 (c) and (X) 3 (MC) Then Kx (MC)(X)yi ~ (MC)2yi KA (M.X)(C) (0)2 01” KX (Mo) 3 (C) ——2‘ (3A-17) K Y _ A i and (MC.X) = KX(M.X)(C) = KX(C)2 (3A-18) Substituting for (MC) and (MC.X) from equations (3A-l7) and (3A-18) into Equation (3A-15) gives, K x 2 KAY: (C) + (C) 306 which can be solved for (C), K Y K Y (o) . A i A l“ ,_ (BA—l9) The concentrations of the other major species are approxi— mately, are known. (M.X) 3 (C) (SA—20) K (X) 2 (MC) = (C)‘/ X2 (3A—2l) KAY+ (Mc.x) : xx(c)2 (3A—22) Now the approximate concentrations of the major species To correct the scheme for the minor species, we utilize the exact mass balance equations. Let us define (M) + (XMX) + 2(MXM) = A (3A-23) (M) — (XMX) + (MXM) = A' 307 Subtracting Equation 3A-9 from Equation 3A-E gives, (M) - (XMX) + (MXM) + (MC) - (X) = O (3A-2A) and subtracting Equations 3A-lQ from Equation 3A-o yields, (M) + (XMX) + 2(MXM) + (MX) -(c) = 0 (BA-25) Substituting A and A' from Equations 3A-23 into Equation 3A-2A and 3A-25 gives (X) = (MC) + A' (3A-26) and (MX) = (C) (3A-27) I D if we define (C)' = (C) A/2 (3A-28) and (MC)' = (MC) + A/2 (3A-29) Equations 3A-26 and 3A—27 give, (X) = (MC)' + A'/2 (BA-30) 308 and (MX) = (C)' - A/2 (BA—31) Then, K 1 I MA = new = Kxnc')new — (a/2)2] (3A—42) The new concentrations of (C) can be used in Equation 3A—27 to give (MX) = (c)new — A (3A—u3) Then the concentrations of the other species, new’ (XMX)new’ and (MXM)new ’ can be obtained from equilibrium constants (Equations 3A—2 to 3A—6) as follows, 311 ( ) (MCX)new ( A“) X = 3A- new 2 KA< ll (X) + (MCX) + (MC2X) (3B-U) i i 3114 FROM NWR TECHNIQUE *i******i***kittititttttiittitttt*tti*tiiitttiittttittt MULTIPLE DATA SET itattittttiiiaittttttitiitttt*tttii*****i*t**i*iitttttt 101 COMPLEX FORMQTION * i o I LoK E0: A64 INFT“. 2LT DA 9F c RH4N3 AKTI: o COS-KR 4. DANNQQ, T DEIEOQF DDRNKYGTH RRCNQDLRI AAHNpAWWt:AK CC BRLBPQ SQU$9:Do C”..CaLTLM.nD.firb Amounn “firth. Tr... ..Urt , PJPHRHFMLL? RP 9 9 AYTT VTDP OIQU NOTL 9599 TPTT PESP 09XU NPQN GAL! RIAT CQVA N060 IIIIU VASE Q QEOT 990A AOTD LVON CAL! T QZT U195 IQPN QTOO ESUC PEFO ATvT T905 dXFC OTQN [UPI P90T AXFC TAQE NIMIV QOTlO .LTTSAY NQUD 9 ) 9n» )1 1‘ 2Y9! Q. 90)) )0 900 42x15r3 ((0“ XTlva VAC‘VIVI OEYTY. )V 'x ' 0 ,)R\I 8,090 I..1..U)5 .230‘ 49(5T (\UT‘P XZSNnU T‘XIL XNPQM’ x 9 9’s, X))GUIC 9002 95 T2rhl\\lt\ P‘SLOT CU‘ArJD. .U ’UV‘O H9)FGTM TITO .IS "I 9 EP0)EC)51 MOSLQVOI' IK90)95(5 [/435)(X9 TT((ZEXX4 DNXO(1RX! EUUOQSEI FOOIQS NOOPYTRONQT89)Q IKXYOSFPO)!5)62ECI T QTJN//10)(01(U6626 U U ITYP 7,819,10911912913) NTtTEMPERATURE) CCCCCCCCCCCCCCCCCCCCCCCC X E Lx pE ML T OF N CH AT 0 TNIC S‘- NTll OS 0 CNFl 00 NC F 0 N0 IND TOIT ATTN MTAA RANT CMRS FRGN CFO NF C 0 R IRIN .IAO EAPI CIEELQLP.T CFPPE/P.NA CFMMIIINOM CESEDiROIR Cn¥ll(tT;1 0 CC : : C )E) CY1C2 CTQNO CITET CVARA CIDED E E 0 ECTJFJ UCC(E( UNKXJCNNSUOSSC N::::NNNN INNCASRS OONRJCOON50T(TOIEERKRIRTU:RIC:T:T RMU99QMVE‘3SISTTPPLNUTUT:AUTCASFS BMOKMYMMVP‘NIN NAAVUTNTNSMTNCMNEN 12,04 DUONXQYOOIOLOXOOOTTOOEOEOTAEOCAORO RSC9OLYCCDFZCDCGCIJNNRCRCIGRCCGCTC 12345 C 2C C C PL :2FF F T 0 T S 5 N C X :4 I E 1 T L o C P 8 N A M 9 C PS 0 2 1 FE CT T I T A EC lTA R SLE 0A T "OP 1 x N OMS Fxc E I: CC” C TA M: N AG T:C 0 RE NC+ NC TN A+X 0 NO TXM IX E! SN N TC CA N FX 0 AL NT Orr-CC I Rh..- 0...... CC MNT TMNCZ T20ANNOO O NTNCIROECISA OVHAH+ TIT-.....C TNT TTSMREAOoRIT ASH TCRCIT.’ "NCFNNT NEA ROCOEONXDELT ))0 50C CCEEECPL 34 RIF NYN CLRNIE QOYIO NIPOP“PIORD TTPABROTLCIOMACTO AALFglIAA ACOP :A DEA-eATyHCP C.NXT JJHN PARTNcN NONE ((TOT.HONANOICTXB SSNIANRFEGCIoITEQ TTE: SSZ) NNAI OOH( CCDU 00::IIN13A9A IF))L:A:XCMH : :34 2 x..cc: XP AX((CMXMMMML PPUUEE~LEFE7.A CCCCCCCCCCCCCCCCCCCCCCCC 315 ' . ___.'- ...-~- - _ * 'k NMR TECHNIQUE MULTIPLE DATA SET-CONTINUED iii*******************tt*i*t**t*i**t*****i**********tt* *******t**********£*************k********i*ti******t*t* 1.1 COMPLEX FORMATION FROM ”k i C I F C . C C E C L T C F XF C NR I CI C 00 R LH C I T PS C TR M C AI C 0L X C CA I CA E C P N C L C E. 0 II P C EN I 0M M C R0 N 1E 0 C FINA H C C 0 XDC C FFIFEE I C 00.0LRF o C E PIO I C TTLTMA TC FFPFOPT FFC IIIIC.NXOIC HHRH NEE HC LLCLI IMILC AAIA FFOTAC CCNCFOFCACC 11.2.0710 E FTC MMIM TCITMC EETETFCoNEC HfiAHFI IEHC CCCCIHE CCC :: :HSRDN C I) )S UEOUC 56 7 LTRCEC Q9 9LAAI VC TT TACRALRC AA ACIEPAEC DD DIMP.TSC1*S 00o0/U(1)))) 1.0/TREF) (JDATQI) T 7. u 0 TI 0 FMATTO(()G EEOAA1((M) RTJDD‘PPAO 285T/(JJ+XXGO) *Tl-OS((OEE*CA .TSS C((A OHAX *XXTl) )( 0T 0R CG 05 dd dMEMNOEC+AN8MlSTTGt*PE(AP2CE+ (( (EHEOTOCSMO9E(NSSE))(LT+ttALB SS SHCTI::CTAC2T:ONN813/0R00*+oAC*XPPXPN(:E/0¢XR*0(U*+0(CE/UotRPC TT TC:: )ICIG:::RCOO3((XAQooBDH.ECE::M:I:TDQ 33 S:)) 12C::PFFP:CC:UUPHSIQAAH(:E:XXEMTTP(CIO:SAIX4DUD::(CIGA:XE NN NC25 ((CSMMEIMT::A:::(:::::(:X2CCM:XNLL:::(8:::::::(PN:::(:CC: 00 OM(( XXCTAERDEDTIHAXAFABCCHFCMXMMXMMOEEABDFBARABDCEFCCAEEFRMHMX CC CDUU XXCIGTTTTSPPDDC:HIAAAAHIEEEEEEEECDDCCGIHARDDDDCIEECCGIREEE C C C C C C CPXLCFFXLCPKC (EC**2) EMX*EX I*EX*G$M) MX*EM ELPT**2)) )/2 .0 (D *X) GMM OEA 2(6) (**2 +)A* )XP * VflMl‘) MX‘TI XE(L E(-E +)2. )Mic ME*EBOPHQR.X ik(tEE(T(CE((SR)P2DED+T(CES**I XT/TU((L(-L/T(+l**DL_PL‘.L‘RXX l 0 1L 7 ...g 0 TI 0) GM )A \I \l D 6 CPU)‘ .*BT. CABR oDFGA DELT+DELPT+(PX*(DELT**2)) *DB PA*GAM)/PX)*(DELPT**2)) )**2) ) ) X P t C O 2 71‘ Z/ ) 0) TE 5 (T 0‘ (L GT (E )R .C SGT). .SL2N 0+EtC .AD*EB oXoACE‘EDA (DELT**2)) T. P) L2 E* Di -\l\l )TM )LA TEG )LCt O)E—C GECNM )G.CE 0(NE* oTC‘A 0RE(P 00((( 0EA(:W 316 t * FROM NMR TECHNIQUE SET-CONTINUED MULTIPLE DATA tinniifiuirhntti“mit“”1ih”d**fluiihuitfiutt“Mitt 1.1 COMPLEX FORMATION *iitiiiifitifiii******ti******iii*iiifitiiittiiitfittiiiiti i i .T) TM. I...“ EsU DiXM nor... in? 0R 37 )GO TO T D Cu) [A )F R+R TnUT 0’s OS) 0N * I t B ... XEEXnuEivl‘OF o... o/ N-EC)/EC”) o t XXXPLTSIOCZC 2.1M. HEELS *Ichfirr * E (Pr—LE... 0(9t bolt ls .(ttXNT2i7PA( 0T9XMS CXPP:RQODIEABNT:EEE::NOCC2U(TE:(XTNN:P(XTNSNN:(XTSNNNNNNENNNNNNNR EN::NTS554zGACAA:::A&0C::5::LM)EMARIDAEMAITRIDENATRIRIRIMRIRIRIRI :EXMES:((.A::ERHAAATGC:MX:XCEIZTXMUTILTXVTIUTITXMIUTUTUTIUTUTUTUT X:MXR(R:z:MMT:(PTTTLE:XXMCCLD((IERTNS(IERN(TNSIER(TNTNTNCTNTNTNTN MHXMTFTABCAAACFLEEEEMMMVXMMA(FXR OEUEFR OCFECER OFEOEOEOFEOEOEOEO N/TTEFLR r. FprESISFFFsgbnhEI-A.HDTZDODODaUDDC .9 IX.” 1 _§)+(TETA*DXMX)+(ZETA*DMXM) :5)JDAToXX‘l)9EC9£MC¢EMCX9EH1EXQEMX9 EnIC MvGA 1591 ) 1 F. c I 1 D T ”CS 0 t 0M3 T a. *0 8‘1Hi0 ) .r.» 9FTAT 4. G ))9IEG o ) )bTZDBEO 0 1 )6 9 91T‘MG 1 0 . \J .3 9TT 0* +0) E 0 )\I1\I 9TAA0)\I( 0 M0 0))11(ITADO . 5M...1 011((X(AD.U.UI\(D\I- o((XXXXDd((+U*Xo fiLXXXX/XJ‘SSS‘ACE oXX/lX/(STTJCOHMNCPX 9 T/IXHCCSTSSI)PD oLAan GMXHXMMTSNNSELaHATES OEMXMHEESNOO 0(AATCJ 9‘ 6? AM 15 =C ( 0 0 A 1 G.NOPT)ITS UC olu 9(U o 1XX(1)QECQEMCQEMCXQENQEXQEMXQ , Tu. 4r, AA o: 0608 IU 91.] \IAEI 4M1) OAT¢T 1.6 9p 9 9:40 [MIN prpx 9 o EAMXG E ITERLC UIIU 9‘ o UUUT ErtpLE UUUHU FRCRIH FCIRACRW FIRALRCRCIRabfiuCRCRC 1 1 1 5 5 60 7. a. 3 a. 5 no 9 3 I O 3 CC 3 no 2 I 1.. 1. 16 an; . .. - .T—w-r. . - . . *‘-L"-“,..a.a$'-_..a_;":n_ LLJ .. _._ -..‘uhi .A_ _ , __ __. k, _ __ __ 317 *******t*****i****t*******************tt*it**t********t * 1.1 COMPLEX FORMATION FROM NMR TECHNIQUE * * MULTIPLE DATA SET-CONTINUED * *ti*************ki‘kki*i‘k'ki'ki'kii'kt‘kit‘k'k‘ki'k‘k'ktik‘k'ki’iitiit RETURN 12 CONTINUE RETURN 13 CONTINUE RETURN END 7 9 CARD ****************t**t*i**i****ttii**t*t***********t***** CONTxOL CARD TITLE CARD MCRT ARRAY CARD NCST ARRAY CARD CDNSTS ARRAY CARDS IRX ARRAY CARD ISNIN ARRAY CARD IRITIAL ESTIMATE CARD DATA CARDS ******************iii******************************tttt BLANK CARD 6 7 E 9 CARD 318 ct = (C) + (MCX) + 2(MC2) + 2(MC2X) (3B—5) Equations 3B-3 and BB—A provide, (X) = (MC2) (BB-6) Subtracting Equation 38-3 from Equation 3B-5 gives, 2U II (c) + (MC2) + (MCZX) (C) + (X) + (33-7) in Wthh R = Ct - Mt' Equilibrium 3B-l gives, (MC2X) Kx2:=(Mc,X)(cj (38-8) Substituting for (MCX) from Equation 3B-7 and rearranging it gives (MC2X) (C) =KXEEMt-(X)-(MC2X)] (3B-9) If we substitute for (C) from Equation 38-9 into Equation 3B—7 we obtain, (MC2X) gdgwt—