a I}... .. 2. 3.6,:frl spa ( :11 r... f... y .51 oer...“ Ma“ .,y.. 42313:...er 6. lo I: 0’ [at 44 ! v 1‘. ‘ , . 1... v7 2:3? .r. ,. :r..1.n...$...v.?finn.r¢r.?. 51.1119. IIIIII3IIIIIII1III2IIIIIIIIIIIIIOIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIHI ’74 b l U :3 “1 / >4 W 00579 8065 LIBRARY Michigan State University This is to certify that the dissertation entitled Spectroscopic Studies of Ionic Solvation and Complexation in N-Methylformamide presented by Lee—Lin Soong has been accepted towards fulfillment of the requirements for Ph.D. degreein Chemistry flaw/WW // 3%, W MS U i: an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE MSU Is An Affirmative Action/Equal Opportunity Institution _————_~_ _ . 7 e ,7,_ _ SPECTROSCOPIC STUDIES OF IONIC SOLVATION AND COMPLEXATION IN N-METHYLFORMAMIDE BY Lee-Lin Soong A DISSERTATION Submitted to Michigan State University in partial fullfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Dpeartment of Chemistry 1988 ABSTRACT . SPECTROSCOPIC STUDIES OF IONIC SOLVATION AND COMPLEXATION'IN N-METHYLFORMAMIDE BY Lee-Lin Soong Lithium-7 chemical shifts in NMF solution are virtually concentration and counter-ion independent, presumably due to a low probability of direct ionic encounter --- an impenetrable solvation shell. Free cations and solvent- separated ion pairs are the likely components for lithium salts in NMF. For the sodium, potassium and cesium salts studied, the chemical shifts are dependent on the anion and change linearly with concentration. The order of increasing shielding shift for cesium salts is: 2-. I” < Br' < c1' < co < son" 3 ' ._ < F' < N03- < c104' < TPB-.- The observed behavior can be accounted for by the formation of collisional ion pairs. Thallium-ZOS chemical shifts change nonlinearly with concentrations of thallium nitrate and perchlorate. This is believed to reflect ion-ion interactions. Ion pair formation constants for these two 1 - salts are 2.6 i 0.4 M- for 111%., and 1.7 1: 0.5 14-1 for I \— _II —II. I—ulll Lee-Lin Soong T1C10 4. Solutions of LiSCN, NaSCN, CsSCN and TIAASCN (tetraisoamylammonium thiocyanate) in NMF were studied in the mid-ir region. The peak position of the (EN stretching vibration for all of the salts studied was 2057 j; l cut-1: ie. the peak frequency of the v1 mode of SCN- in NMF is independent of the nature of the cation and the solute concentration. This suggests the absence of contact ion pairs in these solutions. Curve fitting analysis of these broad bands reveals three bands, at 2045.0 cm-1 1 and 2065.5 cm-l. Since the positions of these three , 2054.5 cm" components are cation independent, the multiplet structure must result from interactions between the thiocyanate anions and solvent molecules. Definitive assignments cannot be provided. Studies of complexation reactions between cesium ions and crown ethers and cryptands were also carried out. The stabilities of Cs+ complexes with crown ethers and cryptands decrease by 18C6 > DBZ7C9 z DBZ4C8 3 C221 2: C222 > DC18C6 > DA18C6£ DBlSCS, DBZlC7 and c211 do not form complexes with the cesium ion. The formation constants are usually small because of high donicity of NMF. Solvent-ligand interaction is a very important factor in determining the stability of the complex. D8210? must have just the right environment for NMF molecules, so there is almost no complex formed with the cesium ion. ACKNOWLEDGMENTS The author would like to express his sincere gratitude to Dr. Alexander I. Popov and Dr. George E. Leroi for their guidance and encouragement during the course of this study. He also wishes to thank the research group of Dr. Popov and H.-H. Nam for their friendship, encouragement and discussions. Great thanks also go to the NMR group of the department for their maintenance of the NMR spectrometers. . Financial aids from his advisers, the National Science Foundation and the Department of Chemistry, Michigan State University are also acknowledged The author also wishes to thank the strong support of his family as well as his wife’s family. Sophie’s and Steven’s loves and smiles, which work just like the catalyst to the reaction, have cheered him a lot. Finally, he ‘wishes to thank. his wife for her patience, encouragement and good careness of children. iv Chapter Page LIST OF TABLES O O O O 0 O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O x LIST OF FIGURES O O O O O O O O ........ 0 ...... O O O I O O O O O O O I O O O 0 ”iii LIST OF ABBREVIATIONS O O ..... O ..... O O O ...... C O O O I O O O O O O XXiii CHAPTER I - INTRODUCTION................................. 1 A. Introduction.................................... 2 B. Historical Part................................. 3 l. Interactions in Electrolyte Solutions......... 3 1.1 Introduction.............................. 3 1.2 Electrical Conductance.................... 5 1.3' Nuclear Magnetic Resonance (NMR).......... 8 1.4 Infrared and Raman Spectroscopy........... 14 1.5 ESR and Electronic Spectroscopy........... 16 2. Complexation of Metal Ions by - -~ Crown Ethers and Cryptands.................... 18 2.1 Introduction.............................. 18 2.2 Selectivity of Complexation............... 21 2.3 Thermodynamics of Complexation............ 27 3. N-methylformamide............................. 29 3.1 Properties and Structures................. 29 3.2 Solvation................................. 32 V Chapter ‘ Page 3.3 Ion-Ion interactions...................... 38 CHAPTER II - EXPERIMENTAL PART........................... 44 A. Materials.... ...... ............................. 45 l. Salts......................................... 45 2. Solvents...................................... 47 3. Ligands.. ........... .......................... 47 B. Methods......................................... 48 1. Sample Preparation............................ 48 2. NMR Measurements ..... ......................... 48 3. IR Measurements ...... ... ..... ................. 49 C. Data Treatment .............. .................... 51 1. NMR Measurements ............ .................. 51 2. IR Measurements.... ..... ...................... 52 CHAPTER III - IONIC SOLVATION STUDIES IN N-METHYLFORMAMIDE SOLUTIONS................ 54 A. Introduction.................................... 55 B. Ion-Ion Interactions in N-Methylformamide Solutions..................... 56 1; Results..... ...... ............................ 56 2. Discussion.................................... 78 2.1 Infinite Dilution Chemical Shifts.......... 78 2.2 Concentration and Anion Dependence of Chemical Shifts.............. 82 2.2-1 Lithium Salts........................... 82 2.2-2 Sodium, Potassium and Cesium Salts...... 87 vi Chapter Page 2.2-3 Thallium Salts.......................... 94 2.2-4 Temperature Study of the CsI/NMF System...................... 95 2.2-5 Chemical Shift as a Function of Mole Fraction of Cesium Salts in NMF......... 95 2.3 Chemical Shift Theory...................... 99 2.3-1 Overlap Theory..........................102 2.3-2 Probability Theory......................105 2.3-3 H. R. S. Theory.........................109 2.3-4 Relaxation Theory.......................110 '3. Conclusion....................................112 C. Ion-Ion Interactions in N-Methylformamide Mixed Electrolyte Solutions.....................115 1. Results.......................................115 2. Discussion....................................115 D. Ion-Solvent Interactions in N-Methylformamide Solutions.....................131 1. Results and Discussion........................l3l - --1.1 Infrared Measurements......................131 1.1-1 Thiocyanate Salts Studies...............131 1.1-2 Computer Resolution of Thiocyanate Bands.......................136 1.1-3 1% (Volume) N-Methylformamide/ Acetonitrile Mixtures...................147 1.2 NMR Measurements.......................:...152 vii Chapter Page 1.2-l H—l NMR Measurements....................152 1.2-2 N-14, N-lS and Cl-35 NMR Measurements...153 2. Conclusion....................................156 E. Suggestions for Further Studies.................156 CHAPTER IV - COMPLEXATION STUDIES IN N-METHYLFORMAMIDE...158 A. Introduction....................................159 B. Results and Discussion..........................160 1. Complexation of Cesium Ions by 1806 and Its Substituted Analogs ..................l60 1.1 18C6.......................................160 1.2 D318C6.....................................169 1.3 DC18C6.....................................l71 1.4 DAl8C6.....................................l7l 2. Complexation of Cesium Ions by Larger Crown Ethers...........................l76 3. Complexation of Cesium Ions by Cryptands......183 C. Conclusion......................................190 D. Suggestions for Further Studies.................193 APPENDICES APPENDIX I - DETERMINATION OF ION PAIR FORMATION CONSTANTS BY THE NMR TECHNIQUE: DESCRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUBROUTINE EQUATION...194 APPENDIX II- DETERMINATION OF COMPLEX FORMATION CONSTANTS BY THE NMR TECHNIQUE: viii Page DESCRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUBROUTINE EQUATIONS..........201 A. Determination of Formation Constants for l:l Complexes...............................201 B. Determination of Formation Constants for 1:1 and 2:1 Complexes.......................207 REFERENCESOOOOOOOO ..... O.........00.0.00....0000000000000215 ix LIST OF TABLES Table Page 1. Nuclear Properties of Some Elements.............. 9 2. Diameter in K of Some Univalent Cations (in Crystals) and Macrocyclic Polyethers.. ....... 23 3. Physicochemical Properties of Formamide (FA), N-Methylformamide (NMF) and N,N-Dimethylformamide (DMF)...................... 30 4. Wavenumber Shifts (cm-1) of N-H and C80 Stretching Bands in Salt/NMF Solutions (4 M) 34 5. Solubility of Salts in NMF at Different Temperatures (g/100 g Solvent)................... 35 6. Lattice Energies, Heats of Solution and Beats of Solvation of Some Alkali Metal Salts in NMF -(kcal/mole) at 26.9°C............................ 36 7. Radii (r) for Solvated Ions and Solvation Numbers (n) for Various Ions in NMF at 25°C.............. 39 8. Ion Pair Formation Constants of Electrolytes in NMF at 25°C................................... 40 9. References for Multinuclear NMR Measurements and Their Chemical Shifts (6) Relative to Those of X Table. - Page Infinite Dilute Aqueous Solutions of Metal Ions.. 50 10. Lithium-7 Chemical Shifts of LiCl and LiBr in NMF....... ...... ................ 57 11. Lithium-7 Chemical Shifts of LiI and LiSCN in NMF .............. . ....... . . . . . . . 57 12. Lithium-7 Chemical Shifts of LiTPB and LiNO3 in NMF.... ............. .......... 58 13. Lithium-7 Chemical Shifts of LiClO4 and LiAsF6 in NMF......................... 58 14. Lithium-7 Chemical Shifts of LiPi in NMF......... 59 15. Sodium-23 Chemical Shifts of NaCl in NMF......... 59 16. Sodium-23 Chemical Shifts of NaBr in NMF......... 60 17. Sodium-23 Chemical Shifts of NaI in NMF.......... 60 18. Sodium-23 Chemical Shifts of NaTPB in NMF........ 61 19. Sodium-23 Chemical Shifts of NaAsF6 in NMF....... 61 20. Sodium-23 Chemical Shifts of NaPi in NMF......... 62 21. Sodium-23 Chemical Shifts of NaSCN in NMF........ 62 22. Sodium-23 Chemical Shifts of NaClO4 in NMF....... 63 23. -Sodium-23 Chemical Shifts of NaNo3 in NMF........ 63 24. Potassium-39 Chemical Shifts of RI in NMF........ 64 25. Potassium-39 Chemical Shifts of KSCN in NMF...... 64 26. Potassium-39 Chemical Shifts of KNO3 in NMF...... 65 27. Cesium-133 Chemical Shifts of CsF in NMF......... 65 28. Cesium-133 Chemical Shifts of CsCl in NMF........ 66 29. Cesium-133 Chemical Shifts of CsBr in NMF........ 66 xi Table_ 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. Cesium-133 Chemical Cesium-133 Chemical Cesium-133 Chemical Cesium-133 Chemical Cesium-133 Chemical Cesium-133 Chemical Shifts Shifts Shifts Shifts Shifts Shifts of of of of of of Page CSI in NMF......... 67 CSSCN in NMF....... 67 CsClO4 in NMF...... 68 CSNO3 in NMF....... 68 CSTPB in NMF....... 69 Cs2C03 in NMF...... 69 Thallium(I)-205 Chemical Shifts Of T1C1O4 in NMFOOOOOOOOO......OOOOOOOOOOOOOOOOOO 7o Thallium(I)-205 Chemical Shifts Of TlNo in NMFOOOOO..0.0000000000000000COOOOOOOO 7o 3 Parameters of Linear Regression for Pure Sodium and Potassium Salts in NMF (Concentration vs. Sodium-23 and Potassium-39 Chemical Shifts)... 71 Parameters of Linear Regression for Pure Cesium Salts in NMF (Concentration vs. Cesium-133 Chemical ShiftS)0.0.0000000000000000000000000I... 72 Cesium-133 Chemical Shifts of CsI in NMF at 273K, 296K and 341KOOOOOOOOOOOO......OOOOOOOOO 79 Parameters of Linear Regression for C31 in NMF at Three Different Temperatures.............. 79 Infinite Dilution Chemical Shifts (ppm) Relative to Water for the Cations 133 + 205 + 7 Li+, 23 + 39 + Na , K , Cs and T1 in a Variety of Solvents....... 81 Parameters of Linear Regression for Cesium Iodide in NMF with Different Values of the Ion Pair xii Table_ 44. 45. 46. 47. 48. 49. 50. 51. 52. Formation Constant (K) and Cs-133 Chemical Shift Difference between Free Cesium Ions and Ion Pairs (55)....................................... Parameters of Linear Regression for Pure Cesium Salts in NMF (Mole Fraction vs. Cesium-133 Chemical Shifts)................................. Comparisons of Mole Fraction Shifts (50) with Crystalline Chemical Shifts (61) for Cesium Halides and Proportionality Constants (m) in Equation 3.5..................................... Concentration Shifts of Cesium Salts in Water and NMF................................. Concentration Shifts of CsTPB in DMF, DMSO, EC and NMF....................................... Concentration Shifts of C51 in MeCN, H 0, 2 EC and NMF....................................... Linear Regression Parameters for Cesium Salts in Water and NMF Using H. R. S. Theory's Empirical .Formula.......................................... Summary of Ion-Association Studies in NMF, Via Different Techniques............................. Cesium-133 Chemical Shifts of NMF Solutions of Mixed CsCl/O.10 M CsI and LiC1/0.10 M CsI........ Cesium-133 Chemical Shifts of NMF Solutions of Mixed Neel/0.10 M CsI and NaTPB/0.10 M CsI....... xiii Page 92 98 100 107 108 108 111 113 116 116 Table. Page 53. Cesium-133 Chemical Shifts of NMF Solutions of Mixed LiTPB/0.10 M CsI and CsTPB/0.1O M CsI..... 117 54. Cesium-133 Chemical Shifts of NMF Solutions of Mixed LiC104/0.10 M C51 and NaClO4/0.10 M CsI.... 117 55. Cesium-133 Chemical Shifts of NMF Solutions of Mixed esc1o4/o.1o M CsI and esno3/o.1o M CsI..... 113 56. Cesium-133 Chemical Shifts of NMF Solutions of Mixed LiNO3/0.10 M CsI and NaN03/0.10 M CsI...... 118 57. Cesium-133 Chemical Shifts of NMF Solutions of Mixed LiBr/0.10 M C31 and NaBr/0.10 M CsI........ 119 58. Cesium-133 Chemical Shifts of NMF Solutions of Mixed CsBr/0.10 M CsI and LiSCN/0.10 M CsI....... 119 59. Cesium-133 Chemical Shifts of NMF Solutions of Mixed NaSCN/0.10 M CsI and CsSCN/0.10 M CsI...... 120 60. Cesium-133 Chemical Shifts of NMF Solutions of Mixed NaI/0.10 M CsI and KI/0.10 M CsI........... 120 61. Cesium-133 Chemical Shifts of NMF Solutions of Mixed TPAI/0.10 M CsI............................ 121 62. .Cesium-133 Chemical Shifts of NMF Solutions of Mixed CsSCN/0.2434 M CsI and CsSCN/0.4816 M CsI.............................................. 121 63. Parameters of Linear Regression for Mixed Salt Systems in NMF................................... 127 64. Molar Molecular Shift (ppm/M) for Mixed Electrolyte Systems in NMF....................... 129 xiv Table. Page 65. Peak Positions ‘V(cm-1) and Half-height Width (HI/2) (cm-1) of C:N Stretching Bands of Four Thiocyanate Salts in NMF......................... 132 66. Curve Fitting Results of Thiocyanate Salts in NMF ....... .............................. 138 67. Curve Fitting Results of Thiocyanate Salts in NMF at Different Temperatures.................... 139 68. Wavenumbers (cm-1) of N-H, C=0 and C-N Stretching Bands in 1% (volume) NMF/MeCN Mixtures........... 148 69. Wavenumber Shifts (cm-1) of N-H, C=0 and C-N Stretching Bands of NMF in 1% (volume) NMF/MeCN Mixtures......................................... 150 70. Proton Chemical Shifts of NMF and Its Solutions................................ 152 71. N-14 Chemical Shifts of Thiocyanate and Nitrate Salts in NMF..................................... 154 72. N-15 Chemical Shifts of LiSCN-15 in NMF.......... 155 73. Cl-35 Chemical Shifts of Perchlorate Salts in NMF......................... 155 74. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (18C6)/(CsI) in NMF................... 161 75. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (18C6)/(CsI) in NMF................... 162 76. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (18C6)/(CsTPB) in NMF................. 163 XV Table_ ~ Page 77. Logarithms of Formation Constants of 1:1 and 2:1 (18C6/Cs+) Complexes of Cesium Salts with 18C6 in Various Solvents......................... 165 78. Limiting Chemical Shifts of Cs+, Cs+(18C6) and Cs+(18C6)2 in Various Solvents................... 166 79. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (DB18C6)/(CsI) in NMF................. 170 80. Variation of Cesium-133 Chemical Shifts with the (DC18C6)/(CSI) Mole Ratio in NMF................. 172 81. Variation of Cesium-133 Chemical Shifts with the (DA18C6)/(CsI) Mole Ratio in NMF................. 173 82. Logarithms of the Formation Constants of Cesium Complexes with Crown Ethers in Various Solvents.. 175 83. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (D821C7)/(CsI) in NMF................. 177 84. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (DBZ4C8)/(CsI) in NMF................. 178 85. Variation of Cesium-133 Chemical Shifts with the ‘Moie Ratio (DBZ?C9)/(CsI) in NMF................. 178 86. Logarithms of the Formation Constants of Cesium Complexes with Crown Ethers in Various Solvents.. 181 87. Limiting Chemical Shifts of 1:1 Complexes of Cesium ions with DBZlC7, DB24C8 and DBZ7C9 in Various Solvents ........ ......................... 183 88. Variation of Cesium-133 Chemical Shifts with the xvi Table. Page Mole Ratio (C222)/(CsI) in NMF................... 184 89. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (C222)/(CsTPB) in NMF................. 185 90. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (C221)/(CsI) in NMF................... 186 91. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (C211)/(CsI) in NMF................... 186 92. Logarithms of the Formation Constants of Cesium Complexes with Cryptands in Various Solvents.. 189 3 93. C-4 and C-5 Separation in 1 C Spectra of Cryptand C221 in Various Solvents............. 192 xvii LIST OF E G RES Figure Page 1. Infinite Dilution of Sodium-23 Chemical Shifts vs. Gutmann Donor Numbers................ 11 2. Structures of Some Macrocyclic Polyethers........ 19 3. Various Stoichiometries of K+-Crown Ether Complexes.................................. 20 4. Selectivity of 18-crown-6: log K (Formation Constant) Values for the Reaction of 18-crown-6 with Metal Ions in Water vs. Ratio of Cation Diameter to 18-crown-6 Cavity Diameter......... 24 5. Selectivity of Cryptands: log K (Formation Constant) Values for the Reaction of Several Cryptands with Alkali Metal Ions vs. Cation Radius........................................... 25 6. (a) Cis- and Trans-Forms of NMF (b) Linear Chain Structure of Liquid NMF......... 31 7. (a) A Proposed Structure of the Short-Lived Solvent-Separated Ion Pair for NaBr in NMF (b) Proposed Structures of Some Higher Ion Associates for RbI in NMF.................... 42 8. Plot of Metal (Li-7 and K-39) Chemical Shifts vs. Concentration of Metal Salts in NMF.......... 73 xviii Figure Page 9. Plot of Sodium-23 Chemical Shifts vs. Concentration of Sodium Salts in NMF............. 74 10. Plot of Cesium-133 Chemical Shifts vs. Concentration of Cesium Salts in NMF............. 75 11. Plot of Cesium-133 Chemical Shifts vs. Concentration of Cesium Salts in NMF............. 76 12. Plot of Thallium-205 Chemical Shifts vs. Concentration of Thallium Salts in NMF........... 77 13. Plot of Cesium-133 Chemical Shifts vs. Concentration of C51 in NMF at Three Temperatures.... ........ ......................... 80 14. Sodium-23 Chemical Shifts vs. Gutmann Donor Numbers. Reference - Aq. Na+ at Infinite Dilution... .................. .................... 83 15. Potassium-39 Chemical Shifts vs. Gutmann Donor Numbers. Reference - Aq. K+ at Infinite Dilution........:................................ 84 16. Cesium-133 Chemical Shifts vs. Gutmann Donor .Numbers. Reference - Ag. Cs+ at Infinite Dilution......................................... 85 17. Thallium-205 Chemical Shifts vs. Gutmann Donor Numbers. Reference - Aq. Tl+ at Infinite Dilution............... ..... ..................... 86 18. Plot of Cesium-133 Chemical Shifts vs. Concentration of C51 in NMF at Four Different K xix Figure 19. 20. 21. 22. 23. 24. 25. 26. 27. Page (Ion Pair Formation Constant) Values............. 90 Plot of Cesium-133 Chemical Shifts vs. Concentration of C51 in NMF at Three Different D (Chemical Shift Difference between Free Cesium Ions and Ion Pairs) Values........................................... 91 Plot of Cesium-133 Chemical Shifts vs. Mole Fraction of Cesium Salts in NMF.................. 96 Plot of Cesium-133 Chemical Shifts vs. Mole Fraction of Cesium Salts in NMF..................' 97 Plot of 6 / 51 vs. Mole Fraction of Cesium Salts in NMF........................... 101 Plot of Cesium-133 Chemical Shifts vs. Concentrations of Lithium Salts Added to 0.10 M CsI Solutions in NMF...................... 122 Plot of Cesium-133 Chemical Shifts vs. Concentrations of Sodium Salts Added to 0.10 M CsI Solutions in NMF...................... 123 .Plot of Cesium-133 Chemical Shifts vs. Concentrations of Cesium Salts Added to 0.10 M CsI Solutions in NMF...................... 124 Plot of Cesium-133 Chemical Shifts vs. Concentrations of Iodide Salts Added to 0.10 M CsI Solutions in NMF...................... 125 Plot of Cesium-133 Chemical Shifts vs. XX Figure - Page Concentrations of CsSCN Added to 0.10 M, 0.24 M and 0.48 M CsI Solutions in NMF.......... 126 28. Deconvolution of the GEN Stretching Band of Thiocyanate Salts in NMF......................... 134 29. Deconvolution of GEN Stretching Bands of Lithium Thiocyanate (0.24 M) in NMF at Three Temperatures..................................... 135 30. Resolved Curves of the GEN Stretching Band for Thiocyanate Salts in NMF..................... 140 31. Plot of Relative Intensities of Three GEN Resolved Bands vs. Concentrations of LiSCN in NMF........................................... 141 32. Plot of Relative Intensities of Three GEN Resolved Bands vs. Concentrations of NaSCN in NMF........................................... 142 33. Plot of Relative Intensities of Three GEN Resolved Bands vs. Concentrations of CsSCN in NMF........................................... 143 34. 'Plot of Relative Intensities of Three C:N Resolved Bands vs. Temperatures in 0.24 M LiSCN/NMF Solution............................... 144 35. Plot of Relative Intensities of Three GEN Resolved Bands vs. Temperatures in 0.49 M NaSCN/NMF Solution............................... 145 36. Plot of Relative Intensities of Three GEN xxi Figure Page Resolved Bands vs. Temperatures in 0.40 M CsSCN/NMF Solution...L........................... 146 37. Plot of Cesium-133 Chemical Shifts vs. (18C6)/(Cs+) Mole Ratio in NMF................... 154 38. Plot of Cesium-133 Chemical Shifts vs. (Ligand)/(Cs+) Mole Ratio in NMF................. 174 39. Plot of Cesium-133 Chemical Shifts vs. (Ligand)/(Cs+) Mole Ratio in NMF................. 179 40. Plot of Cesium-133 Chemical Shifts vs. (Ligand)/(Cs+) Mole Ratio in NMF................. 187 41. Carbon-13 NMR Spectrum of C221 in NMF............ 191 xxii 1- ELLIS LiTPB NaTPB CsTPB LiPi NaPi TPAI TIAASCN 2- EQLEEEIE 3 -EC DMF FA THF DMSO HMPA LIS F BBR V ATION Lithium tetraphenylborate Sodium tetraphenylborate Cesium tetraphenylborate Lithium picrate Sodium picrate Tetrapentylammonium iodide Tetraisoamylammonium thiocyanate N-Methylformamide Acetonitrile Acetone Nitromethane Pyridine Propylene carbonate Ethylene carbonate N,N-Dimethylformamide Formamide Tetrahydrofuran Dimethylsulfoxide Hexamethylphosphoramide xxiii 3. LIQANDE 12C4 15C5 18C6 DA18C6 DBlSCG DC18C6 DBZlC7 DB24C8 DBZ7C9 DB30C10 C222 C221 C211 12-Crown-4 15-Crown-5 lB-Crown-G Diaza-lB-Crown-6 Dibenzo-lB-Crown-G Dicyclohexyl-lB-Crown-6 Dibenzo-Zl-Crown-7 Dibenzo-24-Crown—8 Dibenzo-27-Crown-9 Dibenzo-30-Crown-10 Cryptand 222 Cryptand 221 Cryptand 211 xxiv A.I RO CTO Fundamental studies of electrolyte solutions are centered on ion-solvent and ion-ion interactions. A variety of physicochemical techniques have been applied by numerous investigators to study the behavior of electrolytes, primarily in solvents with intermediate and low dielectric constants. Studies of interionic interactions in solvents of high dielectric constant have been sparse since it is generally assumed that in these media electrolytes are completely dissociated. N-Methylformamide (NMF) is a valuable, but little studied solvent which has a high dielectric constant (182.4 at 25°C, 308 at -40°C) and strong solvating ability (Gutmann donor number : 25) . The combination of these two properties would suggest an essentially complete absence of interionic interactions. However, some studies have indicated that there is ion association for certain salts in NMF. These reports led us to a detailed study of the extent and nature of ion-ion and ion-solvent interactions in this solvent, through the use of several spectroscopic techniques. The results of this investigation are presented in Chapter III. Complexation between crown ethers or cryptands and metal ions has not been studied in NMF, so a systematic study of the interaction between monovalent cations and representative macrocyclic complexants in this solvent was also carried out. It will be presented in Chapter IV. B- W 1. WW5 1-1 M92129 There are many different kinds of interactions in electrolyte solutions such as: ion-ion, ion-dipole, ion- quadrupole, ion-induced dipole, dipole-dipole, dipole- induced dipole and dispersion (London) interactions. The sum total of the energies of these electrostatic interactions can be represented as a power series in the reciprocal of the interaction distance r. In this series, the terms range from r-1 for the ion-ion energies (strongest) to r-6 for the dispersion energies (weakest). In addition, there may be specific chemical interactions in electrolyte solutions such as hydrogen bonding, acid-base interactions and charge transfer interactions. Ion-ion interaction is one of the vexing problems in the study of electrolyte solutions. There is no clear fundamental definition of the term "ion pair", and it can be expressed by many adjectives such as "tight", "loose", "contact", "solvent—shared" and "solvent—separated" ion pair. -On a molecular level, distinctions of these expressions might be made: a free ion is surrounded by several shells of solvent molecules: a contact ion pair, well separated from other ions, is composed of two ions held together by electrostatic forces for at least one vibrational period: one solvent molecule is interposed ' between the two oppositely charged ions of a solvent-shared 4 ion pair: each of the two interacting ions retains its inner solvation shell in a solvent-separated ion pair. Electrical conductance measurements, which depend only on free ions and on clusters containing unequal numbers of positive and negative charges, have been used to determine ion pair formation constants since the early 1900s. The experimental data have been treated by several different models. More recently, spectrometric techniques, especially infrared, Raman and NMR measurements, which probe solutions on a molecular basis, have complemented the macroscopic conductance experiments. NMR is sensitive to the local environment of a magnetic nucleus, and the chemical shift of a cation in a contact ion pair will be different from its value in the free ion or in a solvent-mediated ion pair (the latter two are indistinguishable) . In NMR measurements, a single, population-averaged chemical shift is usually observed, which results from fast exchange between the free cation site and the two types of ion pairs. Therefore, it is impossible to determine the relative concentration of each component by NMR methods. It is not possible to compare the ion pair formation constants obtained by NMR techniques with those from conductance measurements , because NMR technique monitors the equilibrium between free ions (or solvent-separated ion pairs) and contact ion pairs, whereas conductance monitors the equilibrium between free ions and all kinds of ion pairs. Vibrational spectroscopy is similarly sensitive to 5 the local environment of an ion: thus a band due to a contact ion pair will be distinguishable from that arising from a free ion or a solvent-mediated ion pair. A detailed review of these techniques will be given in the following sections. 1.2 W Ostwald (1) was the first to derive an equation relating the ion pair formation constant (K) for 1:1 weak electrolytes to the measured equivalent conductance at a given concentration C andAo the equivalent conductance at infinite dilution: K=A2c /I\o(l\o-l\)- (1.1) However, this equation was found to be invalid for solutions of strong electrolytes. Later, an empirical equation for strong electrolytes was found (2): A=Ao - SCI/2. (1.2) This equation is explained well by the Debye-Hfikel-Onsager theory, which shows that the equivalent conductance at low concentration should be a linear function of the square root of concentration. Onsager (3,4) made an important step in the electrostatic theory by attributing the changes in the equivalent conductance with concentration to two long-range 6 effects of ion interactions, known as the relaxation and electrophoretic effects. In the case of 1:1 electrolytes the Onsager limiting law is given by A =Ao - (o*l\° +0*)c1/2. (1.3) 8.2 x 105 / (135)3/2 * where a 82 / (0T) “2 n 13 ll (D:die1ectric constant, T:temperature(°l<) and n:viscosity) The relaxation term (-0*C1/2) results from the perturbation of the applied field by the asymmetry of the ionic 1/2) takes into atmosphere. The electrophoretic term (-B*C account the decrease in the velocity of an ion due to the counter flow of the solvent in the ionic atmosphere. By an extension of the Onsager limiting law, Fuoss and Kraus (5) introduced the ion association constant into the conductance equation: > II a mo - Steal/21 (1.4-1) 7% ll (1-a) / (cozy; (1.4-2) where S = 0*A +B*, V o is the mean ionic activity .... coefficient and d is the degree of dissociation of the ion pair. Fuoss and Onsager (6) revised this equation by adding I higher order terms to the conductance equation: A =I\o - S(CO)1/2 + Ecolog(co) .. 2 - 3/2 _ + cho KMcow: J2(co) (1.5 1) __ 2 2 K — (1-0) / co v+ (1.5-2) _ _ __ 2 where E - Ele E2 b - e / DkT E1 = 2.302602a2b2 / 24c 5:2 = 2.302602abB* / 15c1/2 02 =xNe2c / 125DkT ln v_,_= -b¢/ 2(1+¢a) The coefficients J1 and J2 are solvent dependent and can be expressed as a function of the size parameter a. This equation has been very successful in fitting conductance data of 1:1 electrolytes with 10 < K < 1000 M-l. For K < 10 M"1 an almost horizontal line in the plot of equivalent conductance versus Cl/2 is obtained, which makes the calculation difficult. For K > 1000 14-1 the slope of the conductance plot becomes (very large and is insensitive to changes of the equivalent conductance at infinite dilution. Many other models have been developed to obtain more accurate fits to the conductance data since 1957, when the Fuoss-Onsager equation was introduced. The conductance at infinite dilution can be determined with better accuracy by 8 usingeither the Fuoss-Hsia or the Pitt equation (7,8,9). Justice (10,11) modified the Fuoss-Onsager equation by replacing the ion size parameter, a, with the Bjerrum distance, q. Fuoss (12) proposed. a model in ‘which the solvent-separated ion pair is an intermediate transition state between ‘unpaired ions and. a contact ion. pair. A distance parameter R was defined as the distance from a reference ion beyond which continuum theory may be applied. Ions in the region a 5 r g R were considered as ion pairs. 1.3 MW The importance of NMR to the study of solution effects was first noted as early as 1951 by Arnold and Packard (13) . In the past 37 years there has been a spectacular growth in the literature concerned with this broad topic. In particular, electrolyte solutions have been widely studied and the knowledge of the structure of these solutions has been considerably advanced by the use of this technique. Nuclear properties of some of the elements are listed in Table 1. All alkali nuclei have the spin greater than 1/2 and therefore the nuclei have a quadrupole moment that should considerably broaden the resonance lines. However, with the expection of rubidium the resonance frequencies are measurable: and in the cases of Li-7 and Cs- 133 the natural linewidths are less than 1 Hz. Thus in most cases chemical shifts can be measured quite precisely. Halogen nuclei also show great sensitivities of both the chemical shifts and, especially, the quadrupole relaxation 9 .Table 1. Nuclear Properties of Some Elements Nucleus Taken from Reference 127. NMR Frequency at 14.09 Kilo- gauss (MHz) (’3) Abundance Spin Relative to Natural Nuclear Sensitivity 1H at Constant Field Li-7 Na-23 K-39 Rb-87 Cs-133 C1-35 Br-79 I-127 T1-205 23.315 15.868 2.800 19.630 7.864 4.892 16.202 12.003 34.619 100 93.08 27.2 100 75.4 50.57 100 70.48 0.294 0.0927 0.000508 0.177 0.0474 0.0047 0.079 0.093 0.192 10 of halide ions to ion-solvent and ion-ion interactions. Halogen NMR should be capable of providing detailed insight into a number of structural and dynamic features of electrolyte solutions. Thallium-205 is the third most sensitive spin-1/2 nuclide (14) and its chemical shift is extremely sensitive to solvent and concentration, which make thallium-205 NMR a good probe of the interactions in electrolyte solutions. Chemical Shift Studies - Bloor and Kidd (15,16) and Richards gt a]._,_ (17) showed that in aqueous solutions the resonance frequencies of alkali cations are strongly concentration and anion dependent. The variation in cationic chemical shifts as a function of concentration and of counter ions is undoubtedly due to cation-anion interaction. Richards gt 1]”, (18,19) postulated three types of ionic interactions which may contribute to the chemical shifts: 1) short range repulsive overlap of the ionic wave functions during collisions: 2) electrostatic polarization of cationic wave functions by neighboring ions; and 3) the van der Waals forces. These studies have been extended to a variety of nonaqueous solvents by Popov and co-workers (20). A linear relationship between Gutmann's donor numbers* and Na-23 infinite dilution chemical shifts were obtained (21) (Figure 1). However, such a correlation doesn't exist in the case of Li-7, where paramagnetic and diamagnetic contributions to the chemical shift nearly cancel each other out. Concentration dependent 11 L! \ “[u\ \ n. \ \ a” \ t \ tl- \ 40 ”h \ 3i , \ D i u‘ “ 9 .b \ a 7I- \ " \ t " ‘“ 3- \ 4- \\ 3- \ \ a =- ~ In 7" l \ 0 ' ” «x In 2 2! I .3. O b- .I- an ‘- 8. g.- 3.. “I. Sr- 3.4 Figure 1. Infinite Dilution of Sodium-23 Chemical Shifts vs. Gutmann Donor Numbers. 1. NM, 2. MeCN, 3. ac, 4. Ethyl Acetate, 5. THF, 5. DMF, 7. DMSO, 8. Py, 9. HMPA, 10. Water. Taken from Reference 21. 12 Cs-133 chemical shifts were used to calculate contact ion pair formation constants for some salts (22). Khazaeli gt ‘g1& (23) studied the concentration and temperature dependence of the Cs-133 chemical shifts for different cesium salts and fitted the data using a model involving ion pairs and triple ions. The great discrepancy between the results from ‘the different. methods was ascribed to theoretical and experimental difficulties of the conductance method. F-19 chemical shifts of the hexafluorophosphate ion in a number of solvents have been measured by DeWitte and Popov (24). Small upfield shifts as a function of NaPF6 concentration are indicative of solvent- separated ion pairs. Hinton and Metz (25) measured the T1- 205 chemical shifts as a function of the concentration for TlNO3 and TlClO4 in liquid ammoniam and suggested the presence of free, fully solvated thallium ions, as well as ion pairs and of higher order aggregates as the concentration increased. Analysis of the low concentration *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 solution of a given solvent S and antimony pentachloride in 1,2- dichloroethane solution. 1,2 DCE s + SbCls ------- -> S(SbC15) -AH(Kcal/mol) = Gutmann donor number 13 data between 0 and 30 C’C allowed the determination of TlNO3 ion. pair formation. constants and of thermodynamic parameters (AH and.AS). Quadrupole Relaxation Studies - For the majority of nuclei with spin > 1/2, an important relaxation mechanism, and usually the predominant one, is that of quadrupolar relaxation. There are two major, but rather different relaxation theories. One relaxation model treats the electric field gradients at the relaxing ion as arising from fluctuations in the electric environment of the ion produced by solvent dipoles and counterions - i.e., electrostatic origin (26,27). The other treats the electric field gradients as arising from short-range interactions resulting when an ion collides with a solvent or other solute particle and suffers a distortion in the symmetry of the electron cloud. This mechanism is not electrostatic because the electric field. gradient is propagated. by a transient rearrangement of the observed ion-electron distribution (28). Since the formation of ion pairs in solution clearly affects both mechanisms, an often-used strategy is to study the concentration. dependence of 'the relaxation ‘time and extrapolate to the limit of infinite dilution. A large number of relaxation investigations have been carried out for Li-7, Na-23 and Cs-133 nuclei (29,30,31). For alkali halides (except fluorides) the concentration dependence of the relaxation rates is quite small and the sequence of Cl- l4 , Br__ and I- ions bears no simple relationship to the ion size. A systematic investigation of ion quadrupole relaxation in a broad range of solvents was also made by Weingartner and Hertz (32). From the infinite dilution values of the relaxation rate of the halide ions, they found that halide ion solvation is weak in formamide, acetone and dimethyl sulfoxide, whereas methanol, ethanol and formic acid are characterized by much more effective relaxation and thus stronger salvation. 1-4 W In 1965, Evans and Lo (33) found a broad band in the far infrared spectra (450-50 cm-1 1 ) of tetraalkylammonium salts in benzene, at z 120 cm- for the chlorides and at z 80 cm"1 for the bromides, which could be assigned to the interionic vibration. Edgell and co-workers (34,35) studied alkali solutions in tetrahydrofuran and dimethylsulfoxide. They also observed broad bands whose frequencies varied with the nature of the cation and the anion: therefore these bands were assigned as interionic vibrations. Popov and co-workers (36,37,38) studied numerous salts and solvents and found that in the far infrared region the vibration of a cation in a solvent cage could be observed. The peak position depends on the mass of the cation and on the solvent. In highly solvating solvents the peak position is independent of the anion while in poorly solvating solvents the band frequencies can vary with the anion, - which showed that anions can penetrate into the inner 15 solvation shell of the cation and lead to the formation of contact ion pairs. The above results indicate that depending on the system, one can observe l) interionic vibrations of a cation-anion pair, 2) vibration of the cation in a solvent cage with one or more of the solvent molecules replaced by the anion and 3) vibration of the cation in the solvent cage without replacement by the anion. It is interesting to note that when a cation complexing agent (crown ether or cryptand) is added to the above solutions the vibrational bands of the last two vibrations disappear and a new band appears which is completely independent of the solvent and of the anion and which corresponds to the vibration of the cation in the macrocyclic cage (39,40). Edgell e1; ah (41,42) investigated the effects of cation, anion, solvent and concentration on the far infrared band of NaCo(CO)4 in tetrahydrofuran solution. The additional Raman and IR data for the Co(CO) 4- anion were explained by the existence of different sites for the anion in tetrahydrofuran. It was concluded that in THF NaCo(CO) 4 exists -mainly as solvent-separated and contact ion pairs (43,44,45). Salts with polyatomic anions (9;th N03-, 0104-, SCN-, etc.) , often show distinct bands for the free and the contact ion paired anion. In the special case of a highly hydrogen-bonding solvent with small polarizable molecules, the vibrational modes of a polyatomic anion involved a 16 solvent-shared ion pair may be sufficiently perturbed to permit their distinction from those of the free anion. This has been observed in liquid ammoniam solutions of some thiocyanate salts (46) . Deconvolution of the bands and quantitative measurements of integrated band intensities yield concentration quotients for the ionic association. Several ion pair formation constants have been obtained in this manner by IR and Raman spectroscopy (47) . Recently, a detailed study of the variation in the Raman spectrum of aqueous NaClO4 with changing Na+ concentration and temperature was reported by Miller and Macklin (48) . Based upon resolution and assignment of components in the featureless contours assigned to u (A and u 2 (E) 1 l) vibrations of the T d perchlorate ion, they calculated the equilibrium constant for contact ion association as 0.022 In"1 at 22 oC. 1.5 NW The major chemical difference between the NMR and ESR methods as applied to the study of ionic solvation is that in ESR at least one component must be specifically paramagnetic. Detailed study of ion pairing is possible because of the appearance of hyperfine structures in the ESR spectra of radical anions. The hyperfine splitting is caused by the interaction between the unpaired electrons and nuclei of the diamagnetic cations. Weissman and co- workers (49,50) found by ESR technique that there are associations of benzophenone, ketyl and naphthalene radical 17 anions with sodium cations. ESR also can distinguish different types of ion pairs. For example, Hfifelmann and co- workers (51) observed two sets of metal hyperfine couplings for sodium naphthalenide. Sodium naphthalenide forms contact ion pairs in tetrahydrofuran and solvent-separated ion pairs are formed as tetraglyme is added. The intensity of the line assigned to ‘the solvent-separated ion. pair increases with the increase in tetraglyme concentration. The equilibrium constant between the two kinds of ion pairs was determined. The spectrophotometric study of ion pair formation was first reported by Symons and co-workers (52). They observed marked shifts in the position of the first electronic absorption band of a variety of iodides as a result of changes in the solvent polarity, counterion and temperature. Griffiths and Wijayanayake (53,54) extended this work and made a systematic study of ion pair formation. They found that interactions with tetraalkylammonium ions cause a large shift to low energy for the first electronic absorption band, and interactions with alkali metal cations cause a small high-energy shift. The former shift is a result of direct displacement of solvent from the iodide and the latter is due to the formation of solvent-shared ion pairs. Gilkerson and co- workers (55,56,57) studied the ion association of alkali picrate salts in 2-butanone and 2-propanol. A. new' band assigned to the contact ion pair was found as an excess of 18 alkali cations was added to the alkali picrate solutions. The ion pair formation constants were calculated and agreed with those from the NMR. and electric conductance measurements. They concluded. that for’ different spectroscopic techniques, the extent of ion pair formation should be the same provided that the same concentration ranges are covered and the same expressions for the activity coefficients are used. 2. ngpleggtign of Metal Long by grog; Ethers and 2mm 2-1 1.853.429.2123 Since Pedersen (58-60) and Lehn (61-64) discovered macrocyclic crown ethers and cryptands which can form stable complexes with metal ions, studies of these ligands and their complexes have become a very popular field of research (65). The structure of some crown ethers and cryptands are given in Figure 2. Particularly their complexes with alkali and alkaline earth metal cations can be used as models for investigation of ion transport through membranes in biological systems (66). Complexes of different stoichiometries can be formed between crown ethers and metal ions. For example, potassium cation can form a 1:1 complex with 18-crown-6 (67), a 2:1 (crown/K+) complex with benzo-lS-crown-S (68), a 1:2 complex with dibenzo-24-crown-8 (69) and a 'wrap around' 1:1 complex with dibenzo-Bo-crown-lo (70) (Figure 3). Two kinds of complexes between cesium ions and cryptand-222 can 19 o _ Fo/fi (Si c”; c: ‘5 bk) K/ 12C4 15C5 18C6 (:[gflm’fitfij [:j:€\v”E):D <_(—\_> 1,0»; k¢&\J *zflkdp’m DC18C6 D318C6 DA18C6 new I». .° . Vb 086%?) aiouoejg . (5.101 DBZlC7 DBZ4C8 DBZ7C9 $513 w W W C222 C221 C211 Figure 2. Structures of Some Macrocyclic Polyethers. 20 K+(18C6) K+(DB15C5)2 W (6°00 age 7.} . @Eo:o:o§oj:® ‘06:,900 @ 2+(Deuce) K+(DB30C10) Figure 3. Various Stoichiometries of K+-Crown Ether Complexes. Taken from References 67-70. 21 also be formed (71). The cation is inside the cavity of the cryptand in a inclusive complex and there is no anion and solvent dependence of the cesium-133 chemical shift. On the other hand, the cation can be only partially inside the cavity in an exclusive complex: then the cesium-133 chemical shifts of the complexes are dependent on the anion and the solvent. Different kinds of physicochemical techniques have been used to study the thermodynamic and kinetic parameters of the complexation reaction, such as potentiometry, polarography, conductance, nuclear magnetic resonance spectroscopy, ultraviolet-visible spectroscopy, infrared spectroscopy, calorimetry and liquid-liquid partion. Of the various methods used for such studies, metal NMR (especially alkali nuclei and thallium-205) has been shown to be a very powerful and sensitive technique since the chemical shifts of metal nuclei are very sensitive to the immediate environment of the metal ion in solution. Recently, a competitive NMR technique has been developed to measure formation constants for cations whose NMR properties are not particularly well-suited to conventional NMR study and formation constants which are too large to be measured directly (72). 2.2 e ct v t Co exa 0 One of the most striking characteristics of the macrocyclic polyethers is their ability to form complexes selectively with various cations. The factors affecting the — 22 formation and stability of these complexes include the relative size of the ion and the macrocyclic cavity, the electrical charge of the cation, the nature of the donor atom in the ring, the number of binding sites in the ring, steric hindrance in the ring, the solvent and extent of solvation of the ion and the ligand, the binding sites and the nature of the counterion. The ionic diameters of some cations and the estimated cavity sizes of the holes of selected cyclic polyethers are listed in Table 2. Generally, crown compounds form the most stable complexes with those metal cations whose ionic radius best matches the radius of the cavity formed by the ring upon complexation (58,59). Figure 4 shows how the complex formation constants in the case of l8-crown-6 vary with the ratio of cation to cavity diameter (73). The maximum stability for complexes of 18-crown-6 with alkali and alkaline earth cations occurs at a cation-to-cavity diameter ratio of unity. Large crown ethers are not as selective as 18-crown-6 because of the formation of complexes of variable stoichiometry. In the case of cryptands; there is a much better relationship between the stability of the complexes and the relative sizes of the cation and the cryptand cavity (Figure 5). Each alkali cation is preferentially bound to the cryptand with the proper size (74). Generally, with a similar cation radius, a bivalent ion has a higher complex formation constant than a monovalent 23 Table 2. Diameter in‘K of Some Univalent Cations (in Crystals) and Macrocyclic Polyethers. Taken from Reference 58. Cation Cation Macrocyclic Macrocyclic diameter polyether polyether diameter (A) (R) Li 1.20 12-crown-4 1.2-1.5 Na+ 1.90 15-crown-5 1.7-2.2 K+ 2.55 18-crown-6 2.5-3.2 Rb+ 2.95 21-crown-7 3.4-4.3 Cs+ 3.34 Larger than 21-crown-7 > 4 T1+ 2.80 Cryptand 222” 2.8 Cryptand 221 2.2 Cryptand 211 1.6 24 4.0 -I 8a" 3.5 -I 3.0-4 .0". tog K 21H C j o o k/‘iJ 2-0 " Cavity Diameter were 1.5“ 1 .0« Co’ 0.5 1f r T fl 1 r 1 0.7 1.0 1.3 Die. MW! Oil. Cavity Figure 4. Selectivity of 18-crown-6: log K (Formation Constant) Values for the Reaction of 18-crown-6 with Metal Ions in Water vs. Ratio of Cation Diameter to 18-crown-6 Cavity Diameter. Taken from Reference 73. 25 10.0w Cwlb 9.0« 122.13” 8.0« [2.1.1]b b .{1zfl 7.o~ - - ‘0‘ [1321‘ beK = s.o« I. ‘ I 4.0- ,’ \. I I I I, 3.0- I I . I I 10., I ‘.:_____: o Tr W T W T 7': 100 125 | 1'50 175 i no Mr K’ my c3: Metal Ion Rodin (pm) Figure 5. Selectivity of Cryptands: log K (Formation Constant) Values for the Reaction of Several Cryptands with Alkali Metal Ions vs. Cation Radius. Taken form Reference 74. 26 ion. ,For example, it has been shown that complexes of bivalent cations with dicyclohexyl-lS-crown-é are more stable than those of univalent cations in aqueous solution (75). However, for small cations, monovalent cations form more stable complexes with dicyclohexyl-lB-crown-G than the divalent cations. This is due to the fact that the divalent cation is more solvated. The type and number of donor atoms in the ring are also important factors in determining the stability of the complex. Substitution of sulfur or nitrogen for oxygen in the crown ether ring reduces the affinity of the ligand for alkali and alkaline earth cations because of the reduction of the cavity size and the lower donor abilities of S and N. However, an opposite stability trend is expected when the cation-ligand bond has appreciable covalent character as in the case of Ag+---N or Ag+---S interactions (76). Cram et a1; (77) investigated the effect on complexation by varying the number of donor atoms in the ring without changing the size of the ring. They found that 18-crown-6 is a much better host for tetrabutylammonium cation than 18- crown-5,— which has one less donor atom in the ring. The formation of a complex in solution is not only a competition for the metal ion between the ligand and the solvent molecules, but also a competition for the ligand between the metal ion and the solvent molecules. Therefore, changes in the solvent can produce significant changes in the formation constants of the complexes. Popov and co- 27 workers (78-83) have studied extensively the complexation of alkali cations by macrocyclic ligands in a wide variety of nonaqueous solvents. The results show that the stability of complexes depends not only an the relative size of the ion and macrocyclic cavity, but an the nature of the solvent. Solvent-ion, solvent-ligand and ion-ion interactions are important factors to the stability of the complexes. For solvents with low dielectric constant, ion pair will interfer with the complexation reaction. The crown-solvent interaction is an important factor which affects the complexation reaction. Formation constants of acetonitrile and nitromethane complexes with 18-crawn-6 have been determined by Mosier-Bass and Popov (84) . The results show clearly that these ligand-solvent interactions can affect the complexation to a great extent. 2.3 111W Kauffman et a1; (85) have discussed the contribution of enthalpy and entropy to the complexation reaction. They postulated that the enthalpy of formation of the complex is influenced by: 1) the replacement of the first salvation shell of the cation by the ligand, 2) the change in the interaction with solvent molecules outside the complex as compared to those outside the first salvation shell of the cation, 3) the change in inter-binding site repulsions, 4) the change in ligand salvation enthalpy upon complexation and 5) the steric deformation of the ligand by the cation. - They also attributed the changes of the entropy of the 28 cryptand complex to: 1) desalvation of the cation, 2) release of solvent bound to the ligand, 3) changes in ligand. internal entropy caused by orientation, rigidification and conformational changes, 4) formation of a single complex from two species, and 5) salvation of the complex. The equilibrium constant for the complexation of cesium ions by cryptand-222 has been studied as a function of temperature in several nonaqueous solvents (71). The values of enthalpy and entropy showed that the complex is enthalpy stablized but entropy destabilized. Popov and co-workers (79,82) have studied the complexation of the cesium ion with large crown ethers in nonaqueous solvents. In all cases the complexation was enthalpy stabilized but entropy destablized. They assumed that the decrease in entropy upon complexation is related to a change in the conformational entrapy of the ligand, although it is not the only factor governing the change in entropy of complexation. There have been many thermodynamic studies to analyze the origins of the macrocyclic effect (63) and cryptate effect.(86), and different conclusions were drawn. Hancock and Martell (87) , in their recent review paper, concluded that for both effects considerable stabilization is derived from the greater basicity induced in donor atoms as ethylene bridges are added. The important factors are: 1) desalvation effects, where steric constraints to salvation of the donor atom in the free ligand lead to increase 29 complex stability, 2) enforced dipole-dipole repulsion in the ligand, which is relieved upon complex formation, and 3) structural preorganization of the ligand such that the donor atoms in the free ligand are already correctly oriented for complex formation. 3. - a d 3.1 Wm N-Methylformamide (NMF) is a methyl-substituted derivative of formamide. Many of its physicochemical properties show intermediate values between those of formamide (FA) and N,N-dimethylformamide (DMF). Comparisons of some properties are made in Table 3 (88). The rotation about the C-N bond of the NMF molecule is very severely hindered (89) by an activation barrier of about 59 kJ/mol, and the trans- and cis-conformations (Figure 6.a) both can exist in the NMF system (The distinction between the cis- and trans-forms is based on the ~C0-NH- bond in peptides). Vibrational spectroscopic studies in the liquid phase (90,91) and in carbon tetrachloride solutions (92) showed that the trans-isomer was predominant in the liquid and in the solution phase. These results agreed with conclusions drawn from gas phase measurements obtained by infrared (93) , microwave spectroscopy (94) and the electron diffraction method (95). NMR spectra obtained in the liquid phase (96) and dipole moment measurements of NMF in dioxane solutions (97) indicated that 90-92% of the NMF molecules were present in the trans-farm. Sugawara e3; 3],. (98) also 30 Table.3. Physicochemical Properties of Formamide (FA), N-methylformamide (NMF) and N,N-dimethylformamide (DMF). Taken from Reference 88. FA Molecular weight 45.041 Boiling point(°C) 210.5 Melting point(°C) 2.55 Density(g/cm3,25°C) 1.12918 Refractive index(25°C) 1.44582 Viscosity(cp,25°C) 3.302 Dielectric constant 111.0 (20°C) pKa(in water,20°C) -0.48 Dipole Moment(Debye) 3.37 Donor number ' 24* Acceptor number 39.8* 59.058 180-185 -3.8 0.9988 1.4300 1.55 182.4 (25°C) -0.04 3.85 24** 32.1* 73.095 153.0 -50.43 0.94397 1.42817 0.802 35.71 (25°C) .000]- * From "The Donor-Acceptor Approach to Molecular Interactions" V. Gutmann, Plenum Press, New York (1978). ** From Yu.M. Kessler, A.I. Mishuatin and A.I. Podkovyrin, £1 59181 989ml. 5. 111 (1977)- 31 (a) H ,H H\ 40 \N"C /~-C\ H—c *0 H—o H / H/it '1 H (trans) (cis) (b) ._. ‘ 4: x - - bondin .. CH NH(CH3) hydrogen 9 Figure 6. (a) Cis- and Trans-forms of NMF. (b) Linear Chain Structure of Liquid NMF. Taken from Reference 99. 32 determined by theoretical calculation that the trans-form is more stable than the cis-form. A linear and flexible chain structure -(Figure 6.b) was proposed for the liquid structure of NMF on the basis of scattered intensity data by the X-ray diffraction method (99). The average degree of association of NMF has been calculated from the dispersions of the dielectric constants by Durov (100). The values varied from 5.3 to 2.8 at temperatures 273 K - 393 K. It is noted that the dielectric constant of NMF is extremely large (182.4 at 25°C) (101), 40% larger than that of formamide, although the dipole moment of NMF is only 15% larger. Such a high dielectric constant can be explained as the result of the fact that NMF molecules are chainwise associated by hydrogen bonding with the single amino hydrogen such that adjacent molecular dipole moments are nearly parallel (102) . Moreover, cyclic hydrogen-bonded dimers, which can lead to a decrease of the dielectric constant, are less readily formed in NMF than in formamide (103) . Thus NMF has a inuch larger dielectric constant than that of formamide. However, the dielectric constant could be decreased as salts are added. to NMF. For example, in 0.7 M NaI/NMF solution the dielectric constant is 100 at 25 °C. 3.2 8211511233 Salvation studies in the NMF-solution system have received some previous attention . From viscos ity measurements on some common salt solutions in NMF, Rastogi 33 (104,105) concluded that the salvation of the cations is through electrostatic bonds and that of the anions occurs through H-bonds: the tetraalkylammonium ions are not solvated due to their large size and low surface charge density. Bukowaka (106) did Raman and infrared studies on LiClO4 and NaC1O4 in NMF solutions. Two possible salvation models were suggested: 1) cation-solvent interaction occurs through the oxygen and (or) the nitrogen atoms of the NMF; 2) very stable 1:2 chelate complexes between cations and NMF molecules could be formed. Banner and Jordan (107) studied the N-H stretching band of NMF for some lithium salt/NMF solutions and found that with polyatomic anions the N-H band shifts to higher frequencies (in the order: PF6_ > C104- > NO3-) and to lower frequencies as the anions are halides (in the order: F- > Cl-z Br- > I-) (Table 4). It was concluded that the interaction of the N-H proton with the polyatomic anions of low charge density is weaker than with the NMF carbonyl group but it interacts strongly with the halide ions in a manner of hydrogen bonding. Paul e3; al. (108) determined the solubility of some alkali and ammonium halides and thiocyanates in NMF (Table 5) . They found that thiocyanates and iodides are the most soluble of the salts studied and attributed this to the higher polarization of these ions and relatively higher covalent character of these salts than that of the corresponding chlorides and bromides. Heats of soultion, heats of salvation (Table 6), ionic salvation enthalpies 34 Table 4. Wavenumber Shifts (cm-l) of N-H and C20 Stretching Bands in Salt/NMF Solutions (4 M). Taken form Reference 107. N-H C=O Salt Av * Av * NaPF6 -23 -26 LiClO4 -80 -38,-26 LiNO3 -l37 -30 Pure NMF -159 -36 Lil -183 -34 NaI -185 -30 LiBr -213 -32 LiCl -214 -34 CsF(dilute) -180,-260(sh) -33 * Calculated from reference 107, all wavenumbers for both bands are subtracted from those of dilute NMF in CCl4 solutions. 35 _Table 5. Solubility of Salts in NMF at Different KBr KI KSCN NH4C1 NH Br NH I NH SCN Reference 108. 20.61 35.20 5.60 29.77 79.47 50.29 10.25 44.21 80.90 6.59 32.55 38.01 74.35 15.09 21.96 4.75 28.99 87.02 35.62 10.42 45.92 77.27 5.57 32.85 58.53 71.74 Temperatures (g/100g Solvent). Taken from 11.10 46.84 73.07 36 Table 6. Lattice Energies, Heats of Solution and Heats of ' Salvation of Some Alkali Metal Salts in NMF (kcal/mole) at 26.9 0C. Taken from Reference 112. Lattice Heat of Heat of Salt Energy Solution Salvation LiCl -201.20 -13.10 -214.30 LiBr -19l.20 -16.14 -207.34 LiSCN -182.70 -11.80 -194.50 LiClO4 -169.40 -13.80 -183.20 NaCl -185.80 -l.95 -187.75 NaBr -176.60 -4.39 -180.99 NaI -164.50 -7.50 -172.00 NaSCN -163.40 -4.78 -168.18 NaClO4 -153.80 -3.35 -157.15 KCl -168.90 +0.30 -168.60 KBr -l61.40 -0.80 -162.20 KI -151.10 -3.20 -154.30 KSCN -148.70 -1.05 -149.75 KClO4 -139.70 +1.50 -138.20 RbCl -162.80 +0.80 -162.00 RbBr -155.90 ----------- RbI -147.3O +1.50 -145.80 RbClO4 -134.90 +2.70 -132.20 CsCl -155.20 +0.55 -154.65 CsBr -149.30 ----------- CsI -140.30 +1.00 -139.30 CsClO4 -128.6O +3.10 ~125.50 37 and ionic entropies were measured by Weeda and Somsen (109,110), Criss g al- (24) and 0111 sf g. (112) from thermochemical studies on some ZNMF salt solutions. They found that NMF has a tendency to approach nearer the anions than the cations. It is known that anions are more solvated than cations in protonic solvents. Emsley gt al. (113) studied the NMF-fluoride systems, using both ab inito LCAO-MO-SCF calculations and spectroscopic techniques (IR, H-1 and F-19 NMR). From the theoretical calculation they found the amide-fluoride hydrogen bond is the second strongest hydrogen bond known. The most stable conformation for the NMF-fluoride complex is to have F- located trans to the carbonyl group with an N- methyl C-H bond eclipsing the C-N band. In support of theoretical calculations, some IR and NMR studies were carried out. In the infrared studies, a "new" 1600 cm-1 band which shifts from 3465 cm-1 (for monomeric NMF) was found as CsF was added to NMF solution. This band was assigned to the N-H (or H-F) stretching band. H-1 and F-19 NMR studies also showed that there is an unusually strong hydrogen bond between NMF and fluoride. However, Symons at al. (114) disagreed with the conclusion that there is such a strong hydrogen bond in the NMF-fluoride system, on the basis of their IR and NMR studies of chlorides, fluorides and xenons in various solvents. Moreover, they concluded that F-19 chemical shifts are not a good measure of the strengths of hydrogen bonds to fluoride. 38 Solvated radii and salvation numbers (Table 7) for various ions in NMF were also calculated by Paul (115) and Pontani gt g1. (116). The salvation numbers of anions have been found to be higher than that of the cations of comparable sizes. The anions may be arranged in the following order on the basis of their salvation numbers in NMF: C1’ > Br- > I‘ > scn‘ > c104'. 3.3 - o teractions Although it is expected that in solvents of high dielectric constant electrolytes should be completely dissociated, interionic interactions have been studied in NMF solutions by several groups of investigators. Paul gt g1. (115) studied electrical conductance of 21 alkali and alkylammonium salts. The data, were analyzed using Shedlovsky’s modification of the Onsager equation (117), and the ion pair formation constants so determined were very small .(Table 8). It should be noted that when association constants are that small, very different results can be obtained. by using more recently proposed conductance equations such as those of Pitts (8), Fuoss and Hsia (9).or Justice (11). Moreover, it has been generally observed that in alkali family salts salvation depends on the charge density of the cations, and therefore varies in the order Li+ > K+ > Na+ > Cs+== Rb+. Obviously, ion pair formation constants should vary in the opposite direction. In the above-cited paper the order for perchlorates is Cs+ sz+ > Na+ z K+ > Li+. Singh and Gopal (118) also 39 Table 7. Radii (r) for Solvated Ions and Salvation Numbers (n) for Various Ions in NMF at 25°C. Ion rd) a na nb Li+ 5.33 5.45 5.4 Na+ 4.29 3.34 3.3 K+ 4.23 3.13 3.1 Rb+ 4.11 2.83 --- Cs+ 3.99 2.51 2.4 NH4+ 3.33 1.50 --- Cl“ 4.43 3.51 3.5 Br' 4.20 2.85 2.9 I" 4.05 2.41 2.4 SCN“ 4.01 2.33 --- C104‘ 4.05 2.25 2.2 a: from reference 28. b: from reference 29. 40 Table 8. Ion Pair Formation Constants of Electrolytes in NMF at 25°C. Taken from Reference 115. Ian Pair Formation Electrolyte Constant(i0.02) LiCl 1.52 NaCl 0.58 KCl 1.18 RbCl 1.09 CsCl 1.05 LiSCN 1.43 NaSCN 1.05 KSCN 1.43 NH4SCN 1.27 LiClO4 1.52 NaClO4 2.17 KClO4 2.04 RbClO4 3.03 CsClO4 2.86 NH4C104 2.17 MeNH3ClO4 3.70 MezNHZCIO4 3.03 Me3NHC104 4.17 Me4NC104 2.33 Et4NC104 2.50 BU4NClO4 2.27 41 concluded from their electric conductance results that there is cation—anion interaction for some tetraalkylammonium salt solutions in NMF. However, it should be noted that the electric conductance studies of French and Glover (119) show negligible association for NaBr and KBr solutions in NMF and those of Johari and Tewari (120) show that MgSO is completely dissociated in 4 NMF. Thermochemical studies by Criss gt g1. initially seemed to indicate some small degree of ion association in NMF (121). However, their later work (122) withdrew the above interpretation. A model of solvent-separated ion pairs was proposed for NaI in NMF solution from vibrational spectroscopy studies (123). It was suggested that a NMF molecule could be attached to one iodide ion and one (or two) sodium ions at the same time. Thallium-205 NMR chemical shift measurements (124) on r111103, TlClO4 and (CH3)2T1+N03' solutions in NMF showed same concentration dependence of the salt, and the ion pair formation constant for (CH3)2T1+N03- was reported as 11 i 10 M-1, indicating that the magnitude of the error makes it impossible to confirm interionic association. Spin-lattice relaxation measurements (125) on lithium ion and sodium ion by lithium-7 and sodium-23 NMR measurements of LiCl and NaBr in NMF show concentration dependence of T A short- 10 lived structure (Figure 7.a) for a solvent-separated ion pair was proposed. The cation is tetrahedrally coordinated 42 “‘ 900 ° (13) D“! 1 Figure 7. (a) A Proposed Structure of the Short-lived Solvent-separated Ion Pair for NaBr in NMF. (b) Proposed Structures of Some Higher Ion Associates for RbI in NMF. Taken From References 125 and 126. 43 by four solvent molecules. The anion approaches the cation, penetrating into the dimple formed by three equatorial solvent molecules. Furthermore, the same authors, on the basis of studies of RbI in NMF (126) suggested that even higher ion associates (Figure 7.b) could be present in this high dielectric constant solvent. W XPERIMENTA ART 45 1- 3.91321 Lithium Chloride (Fisher), lithium bromide (Fisher), lithium perchlorate (Alfa); lithium nitrate (Mallinckrodt) and lithium hexafluoroarsenate (Alfa) were of reagent grade and were used without further purification, except for drying at 100°C for three days. Lithium iodide (Aldrich) was dissolved in acetone and precipitated by placement in a dry ice bath; the precipitate was dried under vacuum ' at 110°C for five days. Lithium thiocyanate (Alfa) was dried under vacuum while the temperature was gradually increased to 110°C over a five-day period. Lithium tetraphenylborate was prepared by the method of Bhattacharyya, gt g1. (128) and was dried at 80°C under vacuum for four days before use. Lithium picrate was prepared by the reaction of 0.5 mole of lithium carbonate (Mallinckrodt) and 0.5 mole of picric acid (MCB) in ethanol. It was recrystallized from a ‘bezene-acetone 'mixture and dried at 100°C for two days before use. Sodium chloride (Alfa), sodium bromide (MCB), sodium iodide '(MCB), sodium perchlorate (G.F.Smith), sodium nitrate (Baker) and‘ sodium hexafluoroarsenate (Aldrich) were used as obtained, except for drying at 60°C for three days. Sodium tetraphenylborate (Aldrich, Gold Label) and sodium thiocyanate (Baker) were used as obtained, except for drying for three days under vacuum at 45°C. Sodium picrate was synthesized by mixing 0.5 mole of . sodium 46 hydroxide (Aldrich) and picric acid (MCB) in ethanol. It was recrystallized three times from ethanol and dried under vacuum at 70°C for five days before use. Cesium fluoride, cesium chloride, cesium bromide, cesium iodide, cesium perchlorate, cesium nitrate (Alfa) and cesium carbonate (Aldrich) were used as obtained, except for drying at 120°C for two days. Cesium thiocyanate (Pfaltz and Bauer) was recrystallized twice from methanol and then dried at 45°C under vacuum for two days. Cesium tetraphenylborate was prepared by the metastatict reaction as described by Mei (78) and then dried for two days under vacuum at 70°C. Thallium(I) nitrate (Alfa) and thallium perchlorate (K a K) were purified by recrystallization from deionized distilled water and then. dried at 120°C for two days. Potassium iodide (Alfa) and potassium thiocyanate (Fisher) were used as received except for drying at 45°C under vacuum for two days. Tetraisoamylammonium thiocyanate (RSA) and tetrapentylammonium iodide (Fisher) were dried under high vacuum at 60°C for five days. Lithium thiocyanate(N- 15) was -prepared according to the method of Rykowsky and Vanderplas (129), by which they synthesized potassium thiocyanate(N-lS). The crude product was extracted with several portions of acetone and the extract was evaporated on a steam bath. The refined product was dried under vacuum at 110°C for one day. 47 2- 22112222 N-Methylformamide (Aldrich) was refluxed over granulated bariumroxide (Fisher) for one day, followed by fractional distillation 'under reducedf pressure ‘with the ‘middle 50% fraction retained. Acetonitrile (Baker) was refluxed with CaH2 for one week, followed by fractionl distillation with the middle 50% fraction retained. Both solvents were stored in a dry box under dry nitrogen over freshly activated Linde 3-3 molecular sieves. The water content of solvents was determined by using a Varian Aerograph Model 920 gas chromatograph and was found to be always below 70 ppm. Deuterated water (KOR Isotopes) was used as received. 3- 1.125225 12-crown-4 (12C4, Aldrich) and 15-crown-5 (15C5, Aldrich) were fractionally distilled under reduced pressure and dried under vacuum for three days. 18-crown-6 (18C6, Aldrich) was purified by converting it to the crystalline acetonitrile complex (130), filtering the solution and driving off acetonitrife under vacuum for two days at room temperature. Dibenzo-18-crown-6 (D818C6, Parish) was recrystallized twice from benzene and dried under vacuum for three days. Dicyclohexyl-le-crown-6 (DC18C6, Parish) was obtained as a mixture of isomers A and B and dried under vacuum for three days. Diaza-lB-crown-G (DA18C6 or [22], MCB), 21-crown-7 (21C7, Parish) and cryptands C222 (MCB), C221 (MCB) and C211 (MCB) were used as received except for drying under vacuum for two days at room 48 temperature. Dibenzo-Zl-crown-7 (D821C7, Parish), dibenzo- 24-crown-8 (DBZ4C8, Parish) and dibenzo-27-crown-9 (DBZ7C9, Parish) were recrystallized twice from n-heptane and then dried under vacuum for two days at room temperature. Pentaglyme (PG) was synthesized by N. Okoroafor following the procedure of Izatt gt g1. (131). It was dried under vacuum for three days before use. B. METEQQQ 1. 22112122551321.2212: For the NMR and IR concentration studies, the samples were prepared by weighing out the desired amounts of salt into 2 or 5 ml volumetric flasks and diluting to the mark with solvents. For the complexation studies, the samples were prepared by weighing out the various amount of the complexing ligand into 2 or 5 ml volumetric flasks and diluting to the mark with the metal ion solution which had been. previously’ prepared. by dissolving the salt in the desired amount of solvent. 2. EHB_H222!£222222 All nuclear magnetic resonance measurements were carried ‘ out on a Bruker WH-180 multinuclear NMR spectrometer with a field strength of 43.2 kilogauss. At this field lithium-7, sodium-23, potassium-39, cesium-133, thallium-205, nitrogen-14, nitrogen-15 and chlorine-35 resonate at 69.951, 47.61, 8.403, 23.62, 103.88, 13.01, 18.25 and 17.64 MHz, respectively. All solutions were measured in 10 mm (or 20 mm for some 49 potassium salt solutions) od tubes (Wilmad) with a 4 mm od insert (Wilmad) coaxially placed inside the larger tube. The insert contained a chemical shift reference and the lock compound (except for lithium-7 which was run without lock). Table 9 lists the references for lithium-7, sodium- 23, potassium-39, cesium-133, thallium-205, nitrogen-14, nitrogen-15 and chlorine-35 NMR measurements and their chemical shifts relative to those of infinitely dilute aqueous solutions of metal ions. All measurements (except temperature studies) were taken at ambient probe temperature, which is approximately 23°C. Sensitivity enhancement. was achieved by signal averaging and exponential line broadening of the free induction decay (132). Since the signal/noise ratio is proportional to (n)1/2, where n is the number of transients scanned, the signal/noise ratio can be improved by increasing the number of scans. Exponential line broadening is the result of multiplying each point of the free induction decay by a decaying exponent. This convolution process improves the sensitivity and doesn’t introduce any lineshape distortion (133). I 3- 13.322222292222 All infrared measurements were carried out on a Bomem- DA3 FT-IR spectrometer with an RSX operating system and POP- 11 computer. The spectrometer has interferometric optics and utilizes a globar source and a mercury cadmium tellwide detector cooled by liquid nitrogen. The sample port was 50 Table 9. References for Multinuclear NMR Measurements and Their Chemical Shifts (a) Relative to Those of Infinite Dilute Aqueous Solutions of Metal Ions. Nucleus Reference 5(ppm) Li-7 0.1 M LiCl/DZO 0.00 Na-23 0.1 M NaCl/DZO -0.08 K-39 sat'd KNOz/DZO 3.00 Cs-133 0.5 M CsBr/020 8.94 Tl-205 0.1 M T1N03/D20 -l.50 N-l4 1.0 M NH4NO3/DZO ---- 15 N 15 1.0 M NH4N 03/D20 C1-35 0.1 M NaCl/DZO ---- 51 purged with nitrogen. For variable temperature studies the sample was placed in a specially designed chamber through which preheated air flowed. Temperature was varied by changing the heating current and controlled. by a thermistor. The spectra were obtained at a resolution of 2 cm-1. A standard demountable Barnes liquid cell with calcium fluoride salt plates and a 0.05 mm path length was used. In these studies the 4000-1100 cm.1 spectral region was covered. C- W 1. NMR Measurgmegts All chemical shift measurements were corrected for the difference in bulk diamagnetic susceptibility between sample and reference by the equation of Martin, gt g1. (127): 6 + 4x/3 (X - X501) x 10 (2.1) = 5 corr 5 obs ref where xref and X sol are the volumetric susceptibility of the reference and sample solvent respectively: 6 corr and “obs are the corrected and observed chemical shifts. Since the salt concentrations, were always low, the magnetic susceptibility of the solution was taken to be equal to the diamagnetic susceptibility of the solvent as indicated by trampleman and Van Geet (134). All chemical shifts are given versus infinitely dilute aqueous solutions of lithium, sodium, potassium, cesium and 52 thallium ions for Li-7, Na-23, K-39, Cs-133 and Tl-205 measurements (Table 9). For N-14, N-15 and Cl-35 measurements, the chemical shifts were given versus the references (Table 9). TheW chemical shift. was determined (135) by 1) a computer fit (Lorentzian function) of the three ‘most intense points or' 2) manual fit (Lorentzian function) of the band profile. These methods are software options in program NTCFTB which was installed in the WH-180 operating system. The choice between methods 1 and 2 was made based on the width of the band. Method 1 was chosen for all bands for which the line width was less than {2- times the frequency' difference of adjacent. points. Otherwise method {2— was used. The uncertainty of the chemical shift was either 2 times the frequency difference between adjacent points or one tenth of the line width. Downfield chemical shifts were taken to be positive. The formation constants of ion pairs and complexes were calculated by fitting the experimental parameters (total concentrations of metaI ions, chemical shift of free metal ions and the observed chemical shifts) with the nonlinear least-squares program KINFIT (136). Details on the use of this program are given in the Appendices I and II (22). 2- IB.M22§E£2E2R£§ For all measurements the solvent spectra were recorded and then subtracted from those of the solution samples to eliminate interferences from the solvent bands. In the studies of the CN-stretching vibration of various 53 thiocyanate salts, it was found that those bands are composites. They were analyzed by the use of a PC program which was written by H.-H Nam (137). The program includes three parts: 1)deconvolution, 2)second derivative and 3)curve fitting. Details on the use of this program are given in Chapter III. C PT I 0 IO -80 O - TH ORMAMIDE O UT ONS 55 A- IETBQDEQIIQN Electrolyte solutions and ionic equilibria in solutions are important not only in many branches of chemistry, but also in biology and in -many industrial processes. In academic research, the field of electrolyte solutions has been largely neglected in recent years. Yet our knowledge of the structure of electrolyte solutions still remains unclear. For example, the over-all salvation number of an ion is an inexact quantity and becomes largely a function of the experimental technique. In order to unravel this complex puzzle, the various Chemical species present in a given solution should be identified, then their interactions and equilibria could be studied. A considerable amount of work is necessary to obtain a better understanding of electrolyte solutions. Toward this goal, one nonaqueous system (N-methylformamide) has been investigated by a variety of physical techniques --- metal ion NMR and infrared measurements, which have proven to be very powerful tools for the electrolyte solution studies. The purpose of this study is to elucidate the role of the solvent in determining the type of species present in N- methylformamide solutions of metal salts. By carefully extending the earlier fundamental studies of Deverell and Richards, Bloor and Kidd, and Popov and co-workers, it is hoped to have more information about the electrolyte solution structure in this solvent. 56 3- IQN:IQE_INIEBAQIIQN2_IE_N:EEIEXLZQBMAHIDB_§QLEIIQE§ 1- 3222.12: The concentration dependences of the Li-7, Na-23, K-39, Cs—133 and Tl-205 chemical shifts of various metal salts in N-methylformamide were examined. The results are given in Tables 10-37 and Figures 8-12. Lithium-7 chemical shifts show no change within experimental errors as the concentrations of lithium salt solutions are varied for nine lithium salts (chloride, bromide, iodide, thiocyanate, nitrate, tetraphenylborate, picrate, hexafluoroarsenate and perchlorate). For all the sodium salts (chloride, bromide, iodide, thiocyanate, tetraphenylborate, nitrate, picrate, perchlorate and hexafluoroarsenate), potassium salts (iodide, thiocyanate and nitrate) and cesium salts (fluoride, chloride, bromide, iodide, thiocyanate, nitrate, tetraphenylborate, perchlorate and carbonate), the chemical shifts change linearly with the salt concentration. The slopes (concentration shift), intercepts and correlation coefficients (showing the quantitative relationship between two variables, if this value equals -1 or 1, all points are on a line) for each salt are listed in Tables 38 and 39. Thallium-205 chemical shifts change nonlinearly with concentrations of thallium nitrate and perchlorate. This is believed to reflect ion-ion interactions. Ion. pair formation constants for these two thallium salts were determined by the procedure described in the Appendix 57 Table 10. Lithium-7 Chemical Shifts of LiCl and LiBr in NMF. Concentration(M) Chemical Shift LiCl LiBr ($0.01ppm) 0.0127 0.0144 1.02 1.02 0.0425 0.0263 1.02 1.02 0.0802 0.0748 1.02 1.02 0.1286 0.1002 1.02 1.03 0.1557 0.1750 1.03 1.03 0.2371 0.2879 1.03 1.03 0.6346 0.4001 1.03 1.04 1.0061 0.8071 1.07 1.04 Table 11. Lithium-7 Chemical Shifts of LiI and LiSCN in NMF. Concentration(M) Chemical Shift LiI LiSCN (10.01ppm) 0.1149 0.0139 1.02 1.02 0.1328 0.0198 1.02 1.02 0.1621 0.0387 1.03 1.02 0.2576 0.0956 1.03 1.03 0.3284 0.1434 1.04 1.03 0.7354 0.2944 1.05 1.03 58 Table 12. Lithium-7 Chemical Shifts of LiTPB and LiNO3 in NMF. Concentration(M) Chemical Shift LiTPB Lino3 5 (:0 . 01ppm) 0.0045 0.0125 1.02 1.02 0.0112 0.0305 1.02 1.02 0.0173 0.0595 1.01 1.03 0.0307 0.0740 1.02 1.02 0.0464 0.1305 1.01 1.03 0.0565 0.2067 1.02 1.02 0.0627 0.2618 1.02 1.03 Table 13. Lithium-7 Chemical Shifts of LiClO4 and LiAsF6 in NMF. Concentratian(M) Chemical Shift LiClO4 LiAsF 6 a (10 . 01ppm) 0.0174 0.0153 1.02 1.02 0.0310 0.0345 1.02 1.02 0.0472 0.0569 1.02 1.02 0.0503 0.1136 1.02 1.02 0.0775 0.1432 1.01 1.02 0.1147 0.2019 1.02 1.02 0.1659 0.3919 1.02 1.02 59 Table 14. Lithium-7 Chemical Shifts of LiPi in NMF. Concentration(M) 5(10.01ppm) 0.0138 1.02 0.0289 1.02 0.0694 1.02 0.1639 1.02 0.2278 1.02 0.2439 1.02 0.2662 1.02 Table 15. Sodium-23 Chemical Shifts of NaCl in NMF. Concentration(M) 5(ppm) Linewidth(Hz) 0.0145 -4.59 42 0.0253 ' -4.70 43 0.0408 -4.59 43 0.0859 -4.68 44 0.0942 -4.55 45 0.1787 -4.57 48 0.2245 -4.55 49 0.2572 -4.55 51 0.2786 -4.65 51 60 Table 16. Sodium-23 Chemical Shifts of NaBr in NMF. Concentration(M) 5(ppm) Linewidth(Hz) 0.0387 -4.66 44 0.1094 -4.65 49 0.2085 -4.62 50 0.2749 -4.61 51 0.4894 -4.56 55 0.6812 -4.51 58 0.9640 -4.39 63 Table 17. Sodium-23 Chemical Shifts of NaI in NMF. Concentration (M) 5 (ppm) Linewidth (Hz) 0.0379 -4.65 44 0.1166 -4.63 46 0.1759 -4.58 48 0.3445 -4.52 51 0.5544 -4.47 55 0.8027 -4.33 61 61 Table 18. Sodium-23 Chemical Shifts of NaTPB in NMF. Concentration(M) 6(ppm) Linewidth(Hz) 0.0199 -4.69 42 0.0427 -4.69 43 0.1008 -4.70 46 0.1774 -4.71 49 0.1823 -4.71 52 0.2059 -4.72 54 Table 19. Sodium-23 Chemical Shifts of NaAsF6 in NMF. Concentration (M) 5 (ppm) Linewidth (Hz) 0.0229 -4.59 42 0.0337 -4.69 43 0.0538 -4.69 43 0.0661 -4.70 44 0.0799 -4.70 45 62 Table 20. Sodium-23 Chemical Shifts of NaPi in NMF. Concentration(M) 6(ppm) Linewidth(Hz) 0.0265 -4.69 41 0.0611 -4.72 49 0.0681 -4.73 50 0.1846 -4.76 60 0.2752 -4.77 70 0.3318 -4.79 76 Table 21. Sodium-23 Chemical Shifts of NaSCN in NMF. Concentration(M) 5(ppm) Linewidth(Hz) 0.0477 -4.69 40 0.1789 -4.67 46 0.1998 -4.66 46 0.2967 -4.65 48 0.3211 -4.53 49 0.4225 -4.62 50 0.5071 -4.60 50 0.8660 -4.58 52 63 Table 22. Sodium-23 Chemical Shifts of NaClo4 in NMF. Concentration (M) 5 (ppm) Linewidth (Hz) 0.0143 -4.69 42 0.0321 -4.70 43 0.0422 -4.69 43 0.1208 -4.70 46 0.2055 -4.70 48 0.4276 -4.71 49 0.5266 -4.71 50 Table 23. Sodium-23 Chemical Shifts of NaNO3 in NMF. - Concentration(M) 6(ppm) Linewidth(Hz) 0.0033 -4.69 43 0.0245 -4.69 43 0.0367 -4.68 44 0.0698 -4.69 45 0.1577 -4.67 46 0.4542 -4.67 49 0.6063 -4.67 51 0.7626 -4.67 53 0.8689 -4.68 56 64 Table 24. Potassium-39 Chemical Shifts of KI in NMF. Concentration(M) 6(ppm) Linewidth(Hz) 0.0258 -2.72 51 0.0559 -2.65 52 0.1567 -2.39 52 0.2220 -2.31 54 0.3642 -1.89 56 0.4354 -l.52 60 0.6183 -1.27 61 0.8066 -0.91 63 Table 25. Potassium-39 Chemical Shifts of KSCN in NMF. Concentration(M) 5(ppm) Linewidth(Hz) 0.0356 -2.74 52 0.0784 -2.73 53 0.1235 -2.69. 55 0.1642 -2.62 57 0.2498 -2.55 72 0.5262 -2.33 78 0.7500 -2.13 82 1.1012 -1.77 87 65 Table 26. Potassium-39 Chemical Shifts of KNo3 in NMF. Concentration(M) 5(ppm) Linewidth(Hz) 0.0498 -2.78 52 0.1883 -2.80 55 0.2569 -2.85 67 0.4973 -2.91 74 Table 27. Cesium-133 Chemical Shifts of CsF in NMF. Concentration(M) Mole Fraction 5(10.08ppm) 0.0296 0.00175 -1.76 0.0428 0.00253 -1.68 0.0629 0.00371 -1.64 0.1001 0.00589 -1.39 0.1294 0.00760 -1.27 0.1478 0.00867 -1.14 ~0.1636 0.00959 -l.02 0.1906 0.01115 -0.98 0.2344 0.01368 -0.69 0.2525 0.01472 -0.57 66 Table 28. Cesium-133 Chemical Shifts of CsCl in NMF. Concentration(M) Mole Fraction. 6(ifl.08ppm) 0.0006 0.00004 -1.85 0.0131 0.00077 -1.54 0.0306 0.00181 -1.43 0.0463 0.00273 -0.81 0.1155 0.00679 0.43 0.2616 0.01524 3.22 0.2933 0.01706 3.98 0.3431 0.01990 4.96 0.4214 0.02433 6.52 Table 29. Cesium-133 Chemical Shifts of CsBr in NMF. Concentration(M) Mole Fraction 5 (i0.08ppm) 0.0038 0.00022 -l.69 I '0.0118 0.00070 -1.43 0.0219 0.00129 -1.12 0.0465 0.00274 -0.50 0.0980 0.00577 1.15 0.1450 0.00851 2.50 0.2213 0.01293 4.94 67 Table 30. Cesium-133 Chemical Shifts of C51 in NMF. Concentration(M) Mole Fraction 6(10.08ppm) 0.0062 0.00037 -1.64 0.0133 0.00079 -1.43 0.0204 0.00121 -1.23 0.0296 0.00175 -0.61 0.0778 0.00458 0.84 0.1316 0.00773 2.91 0.1555 0.00912 3.74 0.1751 0.01025 4.36 0.2354 0.01374 6.80 0.2434 0.01421 6.84 Table 31. Cesium-133 Chemical Shifts of CsSCN in NMF. Concentration(M) b Mole Fraction 6(10.08ppm) . -0.0149 0.00088 -1.54 0.0199 0.00117 -1.54 0.0455 0.00275 ‘1.33 0.0571 0.00337 ‘1.23 0.0903 0.00531 '0.81 0.1605 0.00941 '0.09 0.2443 0.01419 0.74 Table 32. Concentration(M) 0.0039 0.0157 0.0411 0.1035 0.1506 68 Cesium-133 Chemical Shifts Mole Fraction 0.00023 0.00093 0.00243 0.00609 0.00883 of CsClO4 in NMF. 6(:0.08ppm) -1.85 -1.86 -1.90 -1095 -2.05 Table 33. Cesium-133 Chemical Shifts of CsNO3 in NMF. Concentration(M) Mole Fraction 6(10.08ppm) 0.0092 0.00054 -1.85 0.0180 0.00106 -1.80 0.0398 0.00235 -1.74 0.0611 0.00360 -1.69 0.1119 0.00658 -1.54 °0.1809 0.01059 -1.38 69 Table 34. Cesium-133 Chemical Shifts of CsTPB in NMF. Concentration(M) Mole Fraction 6 (10.08ppm) 0.0043 0.00025 -1.95 0.0056 0.00033 -2.05 0.0113 0.00067 -2.47 0.0361 0.00213 -3.81 0.0749 0.00441 -6.19 0.0940 0.00553 -7.32 Table 35. Cesium-133 Chemical Shifts of Cs CO in NMF. 2 3 Concentration(M) Mole Fraction 5(10.08ppm) 0.0054 0.00032 -1.85 0.0140 0.00083 -1.74 0.0241 0.00142 -1.64 0.0344 0.00203 -1.33 ‘0.0490 0.00289 -1.23 0.0664 0.00391 -1.02 0.0832 0.00490 -0.81 0.0953 0.00561 -0.69 0.1361 0.00799 -0.19 0.1799 0.01053 0.32 0.2105 0.01230 0.84 70 Table 36. Thallium(I)-205 Chemical Shifts of TlClO4 in NMF. Concentration(M) 6(:0.03ppm) 0.0041 159.58 0.0072 159.03 0.0184 158.13 0.0402 156.45 0.0868 153.10 0.1432 149.82 0.2252 145.46 0.3045 141.82 0.3697 139.90 0.4728 136.41 Table 37. Thallium(I)-205 Chemical Shifts of TlNO3 in NMF. Concentration(M) 5(i0.03ppm) 0.0044 160.48 0.0155 160.00 0.0284 159.42 0.0607 158.13 0.1056 156.61 0.1591 155.42 0.2093 154.77 71 Table 38. Parameters of Linear Regression for Pure Sodium Sodium-23 Concentration Shift (ppm/M) Intercept (PPm) and Potassium Salts in NMF (Concentration vs. and Potassium-39 Chemical Shifts). Correlation Coefficient NaTPB NaPi NaSCN KI KSCN -4.69 -4068 -4.67 -4069 -4.70 -4.69 -2.77 -2.79 -2.76 0.8835 0.9887 0.9907 -0.9580 0.9547 0.9915 0.9986 -0.9768 72 Table 39. Parameters of Linear Regression for Pure Cesium Salts in NMF (Concentration vs. Cesium-133 Chemical Shifts). Concentration Intercept Correlation Salt Shift (ppm/M) (ppm) Coefficient CsF 5.31 -1.933 0.9970 CsCl 19.83 -1.864 0.9996 CsBr 30.29 -l.824 0.9997 CsI 36.15 -1.877 0.9995 C82C03 12.61 -1.873 0.9969 CsSCN 10.38 -1.782 0.9994 CsNO3 2.67 -1.853 0.9971 CsClO4 -1.34 -1.836 0.9840 CsTPB -59.50 -1.722 0.9998 73 E Q. 3 0— if U) E .9 E-l— m .c Q :E . g ..2_ UCI Kl KSCN KN03 I I I (10 C13 C16 (19 1.2 ODDO Concentration(M) Figure 8. Plot of Metal (Li-7 and K-39) Chemical Shifts vs. Concentration of Metal Salts in NMF. 74 -4.3 Sodium—23 Chemical Shift(ppm) ' 0.0 0.3 0.5 0.9 1.2 Concentration(M) Figure 9. Plot of Sodium—23 Chemical Shifts vs. Concentration of Sodium Salts in NMF. 75 8—- E 51 o. 3 i3 00 6 4- .9 E o .c L) ro 2— pf) | E .2 8 09 C) o 0 cs 0 A CsBr _ D CsCl o O CSF "2 I I I I 1 Oi) 0.1 0J2 On3 C14 C15 Concenhofion(M) Figure 10. Plot of Cesium-133 Chemical Shifts vs. Concentration of Cesium Salts in NMF. 76 O- E Q. :3 . E -2- (fl '5 .9 E m .c L) —4fi I"? PI) | E 3 '6 -6- m x CsZC03 L) o CsSCN A CsN03 Cl CsC|O4 o CsTPB ‘—8 I I I OI) 0.1 C12 013 Concentration(M) Figure 11. Plot of Cesium-133 Chemical Shifts vs. Concentration of Cesium Salts in NMF. 77 170-1 E Q. 3 1609. :E (J) ‘5 .9 ° E CD 150‘ o .c C) 03 O 0 OJ l E .2 140- '5 .c F— o A—A THWD3 o TKNO4 130 , 121 I I 01) 0.1 C12 2 C13 CI4 C15 Concentration(M) Figure 12. Plot of Thallium-205 Chemical Shifts vs. Concentration of Thallium Salts in NMF. 78 I, and the results are 2.6 i 0.4 M-1 1 for TlNo3 and 1.7 i. 0.5 M' for TlClO4. The concentration dependences of Cs-133 chemical shifts of cesium iodide at three temperatures (273K, 296K and 341K) were examined. The results are given in Table 40. For all three temperatures, the cesium-133 chemical shifts also change linearly (Figure 13). The slopes, intercepts and correlation coefficients are shown in Table 41. 2. Qiscussgan 2.1 Infinite Dilution Chemica; Shigtg Infinite dilution chemical shifts for alkali metal and thallium ions in NMF are given in Table 42 along with those obtained in a variety of other solvents. This result provides the starting point for discussion of the factors which may determine salvation-induced chemical shifts. According to expectation, the range of infinite dilution chemical shifts of alkali metal ions increases with the ion size, being (for common solvents) ca. 5-6 ppm for Li+, ca. 15-20 ppm for Na+, ca. 30-40 ppm for K+, and ca. 130 ppm for Cs+. Bloor and Kidd (15) provided good arguments that Na+-23i ~shielding differences between solvents are predominantly due to the paramagnetic term, and this certainly applies for larger cations as well. According to the overlap model of Chemical shifts, the overlap of the ion's outer p orbitals with outer orbitals of neighboring solvent molecules should determine the infinite dilution chemical shift. Attempts have been made to correlate the 79 Table 40. Cesium-133 Chemical Shifts of CsI in NMF at 273K, 296K and 341K. Cesium-133 Chemical Shifts (10.08 ppm) Concentration (M) 273K 296K 341K 0.0062 -5.35 -1.64 2.95 0.0133 -4.92 -1.43 3.28 0.0204 -4.76 -l.23 3.50 0.0296 -4.30 -0.61 3.87 0.0778 -2.70 0.84 5.76 0.1316 -0.79 2.91 7.80 0.1555 0.04 3.74 8.80 0.1751 0.45 4.36 9.57 0.2354 2.73 6.80 12.06 Table 41. Parameters of Linear Regression for CsI in NMF at 5 Three Different Temperatures. Concentration Intercept Correlation Temperature Shift (ppm/M) (ppm) Coefficient 273K 34.59 -5.43 0.9994 296K 36.38 -1.89 0.9994 341K 39.42 2.70 0.9999 80 14— E 10— Q. 3 :5 - U7 6 6- .9 E m - .c C) r0 2% P0 I '1 E .2 8’3 —2- C) _ A 341K 0 296K 6 0 273K ‘ l I I l ”i 01) 0.1 C12 C13 C14 C15 Concentration(M) Figure 13. Plot of Cesium-133 Chemical Shifts vs. Concentration of Cal in NMF at Three Temperatures. 81 Table 42. Infinite Dilution Chemical Shifts (ppm) Relative to Water for the Cations 7Li+, 23Na+, 39K+, 133Cs+ and 205T1+ in a Variety of Solvents. Solvent D.N.* Infinite Dilution Chemical Shift(ppm) Li+ Na+ K+ Cs+ T1+ MeNO2 2.7 -0.36 -15.6 -21.1 -59.8 ---- MeCN 14.1 -2.80 -7 -0.41 32.0 ---- DMSO 29.8 -1.01 -0.11 7.77 68.0 359.7 PC 15.1 -0.61 -9.4 -11.48 -35.2 ----. MeOH 19 -0.54 -3.8 -10.05‘ -45.2 ---- DMF 26.6 0.45 -5.0 -2.77 -0.5 146.8 AC 17.0 1.34 -8.4 -10.48 -26.8 ---- Py 33.1 2.54 1.35 0.82 -31 783.2 Water 18 0 0 0 0 0 NMF 25 1.02 -4.69 -2.77 -1.84 160.8 * Gutmann's donor number Except for NMF data, all data are taken from references 140 and 150. 82 infinite dilution chemical shifts with quantities expressing the electron-donating ability of the solvent. In our group, we have considered the Gutmann solvent donor numbers, which were mentioned in the introduction. There is a rather good linear relation between Na-23 infinite dilution chemical shift and solvent donicity (Figure 14) (21). The correlations are also relatively good for K-39 (138), Cs-133 (139) and Tl-205 (140) (Figures 15,16 and 17). No correlation was found in the case of Li-7 (38,141), because for lithium-7 nucleus the diamagnetic and paramagnetic terms are of the same order of magnitude. The donor number of NMF, which was not determined by Gutmann, can be obtained from Figures 14-17 and they are 24.3, 24, 26.2 and 24 respectively. These numbers are in very good agreement with each other and show that NMF is a solvent with medium solvating ability. 2.2 o oentrat d Anion Do a nos 0 c 8 2-2'1 LEW Li-7 chemical shifts in NMF solution are virtually concentration and counter-ion independent, presumably due to a low probability of direct. ionic encounter --- an impenetrable solvation shell. It is reasonable to say that there is no contact ion pairs formed for lithium salts, because the chemical shifts of a cation in a contact ion Pair (nearest-neighbor interaction) should be different from its value in the free ion or in a solvent-separated 1011 pair. However, the NMR method cannot distinguish 83 '3 \ FN‘ Ur- \ n»- \ \ Ul- \ \ "I. \ ‘. m- ‘ 3' ‘ a a " "‘ 5 " \ a ’b \ .- \ t " \‘i a 3!- \. ..-.--..--..--..‘ 4|- .\ D \ 3'- . \ 2h- \ 3 ' ' 70“ l C ' \ 0- ‘0‘ ' I“ -l- I \\ I -2. ' a‘ -3 L P L l_ 1 L g 1 I 1 O 4 I 12 H N 24 a 33 1‘ O comm Figure 14. Sodium-23 Chemical Shifts vs. Gutmann Donor Numbers. Reference - Aq. Na+ at Infinite Dilution. Taken From Reference 21. 1. NM, 2. MeCN, 3. AC, 4. Ethyl Acetate, 5. THF, 6. DMF, 7. DMSO, 8. Py, 9. HMPA, 10. H20. 84 ‘hnl o 3 3 O 0 b . 15“.. OM... CM“! Sam 3 0 h 6 a Figure 15. Potassium-39 Chemical Shifts vs. Gutmann Donor Numbers. Reference - Ag. K+ at Infinite Dilution. Taken from Reference 138. 85 so ‘ufinoz #- “ ~‘ .MoOH 4O " x 6“” - ~‘HFOf M0 20 " ‘\2 E :- “~“ FG'NHZ a Q 3 O .......-....-.--;‘.DMF .H20 «0 I I I \‘ -20 t I \ n \‘ ’ omen : OP9~ -40 I- . p I I ‘5° ’ 2 DMSO 1 1 L 1 1 1 1 #1 1 2 9 l4 20 26 32 36 DONOR NUMBER Figure 16. Cesium-133 Chemical Shifts vs. Gutmann Donor Numbers. Reference - Aq. Cs+ at Infinite Dilution. Taken from Reference 139. 86 I I I r I r r ‘ I 1 I 2000’ b .0 '1 y- d l500" - 1- .1 1000- ‘ a P ‘ l " d a .. .1 \ 500— 4 .- I- d h _ _ I- 11 : E---------z-§ 94 : I .4 0 ’ ' 7 ‘ I- I " 2 p I «I a - : - III I- q z __ ' b o -800 I I- ' " I- . 4 r l ‘ I- I d I “000 1 11 1 1 11 1 1, 1 1 l 1 7 20 30 40 IO DONOR nausea o 3 Figure 17. Thallium-205 Chemical Shifts vs. Gutmann Donor Numbers. Reference - Ag. Tl+ at Infinite Dilution. Taken from Reference 140. 1. H O, 2. FA, 3. NMF, 4. DMF, 5. DMSO, 6. DNA, 2 7. HMPA, 8. Py, 9. n-Butylamine. 87 between free ions and solvent—separated ion pairs. Weingartner's nuclear magnetic relaxation results (125) showed that LiCl- can form solvent-separated ion pairs in NMF and Perelygin’s infrared results (123) also showed the possibility of the formation of the solvent-separated ion pairs. Thus, free cations and solvent-separated ion pairs are the likely components for lithium salts in NMF. The conclusion does not conflict with the conductance results (115), which showed that ion-association occurs for lithium salts in NMF, because conductance measurements cannot distinguish between contact and solvent-separated ion pairs, whereas NMR is sensitive to the formation of contact ion pairs. 2.2-2 godium. gotassium and Cesium gals: For the sodium, potassium and cesium salts studied, the chemical shifts change linearly with concentration and also depend on anions. The order of increasing shielding shift for cesium salts is: 2- 3 < 3' < N03- < c104" < TPB'. This sequence is the same as those in the KI/HZO (l6) and I- < Br- < c1" < do < SCN- Rb+/K20 (17) systems. Bloor and Kidd related this sequence to the anion basicity and with the nephelauxetic series (163). In the cases of tetraphenylborate, picrate and perchlorate salts, increase in the salt concentration results in an upfield shift of the metal resonance. Whereas in the cases of halide, thiocyanate, carbonate and nitrate sa co te so de th ac E10 an do wh si. we po RU i0: mi: pm of are in SUC 88 salts, a downfield shift of the metal resonance is found as concentration increases. For the bulky, Symmetric tetraphenylborate and perchlorate anions, replacement of a solvent molecule by the anion decreases the electron density around the cation, resulting in upfield shifts of the metal resonance. On the other hand, other anions (except the nitrate ion in the case of potassium nitrate) act as better electron donors than the displaced solvent molecule, and the metal resonance shifts downfield. Akitt (142) suggested that the upfield shift caused by oxyanions and tetrafluoroborate in water arises because these anions do not overlap the cation’s orbitals, but‘displace water, which does have an overlap contribution. This is unlikely, since in NMF downfield shifts for carbonate and nitrate were observed. Then questions arise: what kind of role do the anions play? Perhaps the most obvious possible mechanism is the polarization of the electronic environment of the magnetic nucleus by the strong electrostatic field produced by other ions in its vicinity. If this were an important factor we might expect the smaller, more strongly polarizing ions to produce the largest downfield shifts, so that the shielding of the metal nuclei would show the following anion dependence Cl- < Br- < 1'. However, the observed effects are quite the reverse, and it seems unlikely that changes in the strength of solvation of the ions could account for such a reversal . Electrostatic effects are also 89 unsatisfactory explanations since they are not able to explain the observed upfield shifts of metal resonance in solutions of some nitrate, picrate and tetraphenylborate salts. The interaction must be of another kind. What causes this linear character? Do the salts form contact ion pairs? It is interesting to investigate the possibility in detail by calculating the ion pair formation constant from the equations in Appendix I: a = [-1 + (1 + 4K¥§C>V21(1/A)(:s ), (3.6) where K is a constant, <1/R3> is the expectation value taken over all the outer p electrons of the central ion, A 2 is the sum of the is the average excitation energy and SS squares of appropriate overlap integrals of the p orbitals of the central atom with s orbitals and with p orbitals in both a and a: bonding configurations of the surrounding ions. Deverell and Richards used the Kondo-Yamashita formalism and obtained the chemical shift of an ion in solution as the following: 6 ' K<1/R3>(1/A) [Qion-ionm) + ¢ion-water(c) 103 ‘4’ (0)]: (3.7) ion-water where °ion- (C) and m ion-water(c) are the appropriate ion sums of overlap integrals with all surrounding ions and water molecules at a solute concentration C, and “’ion- waterm) is the sum at infinite dilution. They concluded from the observed linearity of the chemical shift with solute activity that the first term, 41 ion-ion‘c)’ being dependent on ion-ion collisions, is predominant. Since °ion- ion‘c) and °ion-water‘c) are proportional to the activity (or concentration), the slope (activity shift or concentration shift) of the straight line in the chemical shift vs. activity plot is equal to K<1/R3>(l/A) and the intercept. his -K<1/R3>(l/A)¢ion_water(0). ¢1°n_water(0) cannot be obtained, so experimentally the intercept (chemical shift of the metal ion in infinitely dilute aqueous solution) was assumed to be equal to zero. In the NMF system, the same trend of the results was obtained as that in water, and some discussion follows: l)Because of the high value of the dielectric constant, the mean-ionic activity coefficients of various metal salts in NMF solution are all almost close to unity. Thus, in the NMF system the concentration shift is the same as the activity shift. There is a linear relationship between the concentration (or activity) and the chemical shift. This is different from the water system, in which the linear relationship was found only when activity was used. There 104 are two reasons for this. One is that the dielectric constant of NMF is much higher than that of water. The other is that a different concentration range was used in these two systems. For the water system the concentration went up as high as 10 M; in NMF the highest concentration used was around 1 M. 2)The cesium cation has the largest concentration shift in the alkali metal ion series, and Li+ has no concentration shift. This could be due to the strong solvation of the lithium ion, which reduces considerably the probability of ion-ion contact; on the other hand cesium ion is the least solvated, which leads to the largest overlap. For halide salts, iodide has the largest concentration shift because of its largest overlap with the metal ions. 3)For the NMF system, equation (3.7) can be modified into: - 3 b 6 "K<1/R >NMF(1AA)NMF[°ion-ion(c) +‘9ion-NMF(C) 3 . -‘pion;NMF(o)] - K<1/R >water(IAA)water¢ion-water(°)(3'8) In the same way, the intercept (infinite dilution chemical shift) is equal to K<1/R3>NMF ( VA) NMFq’ion-NMF (0) 105 - K<1/R3> (l/A) water water°ion-water(o)' and the slope (concentration shift) is equal to 3 K<1/R >NMF(1/A)NMF' Table 42 shows that sodium-23, potassium-39 and cesium-133 ions have negative values for the infinite dilution chemical shift. In these cases, it might be assumed that . 3 °ion-water(°) 15 larger than °ion-NMF(°) if <1/R > and A are approximately the same for water and NMF systems. 2.3-2 zreeabilisz_1999r1 Covington and Newman (147) wrote a more realistic equation than equation (3.8): 5(C) = K(1/‘A)Wion-ion - °ion-solvent] x P(C)' (3‘9) where 0 10 is the change of overlap when a solvent n-solvent molecule is removed from the close vicinity of an ion, ¢ion- ion is the change in overlap when an ion is placed in the hole left by the solvent molecule, and p is the probability that such an encounter takes place. It is possible that high probability will affect the curvature of the chemical shift vs. concentration plot, whereas large changes in overlap integral will affect the shift magnitude. From equation (3.9), the concentration shift should be equal t° K<1/R3>(1/A)k if ”ion-ion“) -°ion-solvent(cn 106 x P(C) = k. x (concentration). Then three variables will be solvent dependent -A , and k. It might be reasonable 3 to assume that A and are at approximately same order of magnitude for the same metal salts. Then for cesium salts, k(NMF) must be larger than k(water) because the concentration shift in the NMF system is always larger than that in the water system (Table 46). Of course, k is the combination of many factors, such as solvation energy, the closest distance to which the ions can approach, dielectric constant, reorientation of solvent molecules, strength of hydrogen-bond, and solvation number. Not one of them is predominant. For example, dielectric constant is not the major factor leading to the linear change in the chemical shift. If dielectric constant was the major factor, then the concentration shift should be lower for those solvent: with higher dielectric constant values because of the lower probability for ion-ion encounter. However, Tables 47 and 48 show that, generally the higher the dielectric constant is, the higher the concentration shift is for cesium iodide and cesium tetraphenylborate in various solvents. For solvents, with high dielectric constants, the presumption that a dielectric continuum can represent the solvent is no longer tenable. The molecular nature of the solvent and specific ion-solvent interactions must be taken into account. The region in space where dielectric saturation takes place is an appreciable fraction of the space within which ion-ion interactions are considered to happen. 107 Table 46. Concentration Shifts of Cesium Salts in Water and NMF. Concentration Shift (ppm/M) Salt Water* NMF CSCl 5.03 19.83 CsBr 11.15 30.29 CSI 19.64 36.15 CSN03 “2.74 2.67 * Calculated from the data in reference 18. 108 Table 47. Concentration Shifts of CsTPB in DMF, DMSO, EC and NMF. Dielectric Gutmann Concentration Correlation Solvent Constant D.N. Shift (ppm/M) Coefficient * DMF 36.7 26.6 -19.54 -0.9977 DMSO* 45.0 29.8 -11.32 -0.9925 8c** 89.1 16.4 -3s.4o -0.9980 NMF 182.4 25 -59.50 -0.9998 Table 48. Concentration Shifts of C51 in MeCN, H 0, EC 2 and NMF. Dielectric Gutmann Concentration Correlation Solvent Constant D.N. Shift(ppm/M) Coefficient MeCN* 36.0 14.1 1.07 0.9958 *** H20 78.6 18 19.64 0.9947 sc** 89.1 16.4 38.78 0.9979 NMF . 182.4 25 36.67 0.9997 * Lu Lu L. Hsu, Master Thesis, Michigan State University (1976). ** Lee-Lin Soong, Research Report, Summer (1985) and Winter (1986). *** Reference 18. 109 However, the presence of another ion strongly affects the electric field in the region between them. It is then extremely difficult to estimate the effective dielectric constant in the non-spherically symmetric region in between the ions. Non-electrostatic effects: donor-acceptor interactions, hydrogen bonding, etc., are superimposed on the Coulombic interactions. Under such conditions the Bjerrum (148) approach must be abandoned. 2.3-3 5, g. 8. Theory In another paper, Richards and co-workers (18) provided the formula: 5 = P(CHJI '1' 62 + 53]. (3-10) where 51, 5 2 and a 3 are the shift contributions due to ionic encounter, solvent displacement and solvent reorientation respectively. P(C) is the probability of ionic encounter. As before, they neglected the latter two contributions. Then they used Debye-H’ticker potential theory, modified by adding an oscillating potential to take into account the particular nature of the solvent and the tendency for solvent molecules to form well-defined shells around an ion. Empirically, they found that the shift obeys the following relationship: log ‘obs = g x log (concentration) + D, (3.11) 110 where g is a parameter which is anion dependent and is related to the closest distance to which ions can approach. D is a constant. It would be interesting to test this equation with the experimental data from NMF solutions. There is a problem for those systems for which chemical shifts are all (or partly) negative values, because it is not possible to take the logarithm of negative numbers. For the CsCl, CsBr and C51 / NMF systems, only points with positive values of the chemical shift were picked. In the C82C03, 3 negative chemical shift values were converted into positive CsF, CsSCN, CsClO4, CsNO and CsTPB / NMF systems, values by multiplying each by -1. The results are listed in Table 49. It is found that the experimental data for CsCl, CsBr and C31 / NMF systems have better correlation with equation (3.11) than those of the other cesium salt/NMF systems. 2-3-4 W Mishuatin and Kessler (149) hypothesized that the significant solvation symmetry effect should make it possible to perform reliable discriminations between contact ‘ and solvent-separated ion pairs. However, Weingartner (125) opposed the significant electronic contribution to the relaxation process which was suggested by Mishustin and Kessler, and proposed a cation-anion- solvent model for the LiCl and NaBr / NMF systems by analyzing the concentration dependence of 1/T1 in terms of two effects: 1) a direct contribution to q2 (mean. square 111 Table 49. Linear Regression Parameters for Cesium Salts in Water and NMF Using H. R. S. Theory's Empirical Formula. Correlation System g value Coefficient CsCl/NMF 2.173 0.9920 CsBr/NMF 1.785 0.9983 CsI/NMF 1.815 0.9887 CsCl/H20* 0.785 0.9999 CsBr/H20* 0.814 0.9981 CsI/H20* 0.813 0.9985 CsF/NMF -0.469 -0.8957 CsZCO3/NMF -0.537 -0.7874 CsSCN/NMF -0.991 -0.7990 CsClO4/NMF 0.026 0.8780 CsNO3/NMF -0.091 -0.9278 CsTPB/NMF 0.423 0.9850 * take from reference 18. 112 electric field gradient) due to the Coulomb charges of neighboring ions, and 2) an indirect effect caused by distortion of the solvent structure around the reference ions by short duration collisions with counter ions. No contact ion pairs or triple ion aggregates were found, and a short-lived structure for the solvent-separated ion pair was proposed. This result agrees with that from the Li-7 chemical shift measurements, which show that there is no concentration dependence of the Li-7 chemical shift. However, in another recent paper (126) Weingértner proposed that there are various kinds of higher ion aggregates in the RbI/NMF system and suggested that this could be true also in the CsI/NMF system. It is hard to rationlize this predication with the results of the cesium-133 chemical shift measurements. 3- 9292111191193 It seems that there is no single theory which can satisfactorily explain the origin of the chemical shifts for all systems. Overlap theory seems the best, but it requires more complicated and accurate quantum mechanical calculations, which may be powerless if too many assumptions are assumed. This is true especially in the case of NMF, for which the molecular structure is more complicated than that of water. Table 50 summarizes some results on ion-association studies obtained by different techniques. It should be noted that it is questionable whether these experimental 113 Table 50. Summary of Ion-Association Studies in NMF, Via Different Techniques. IR NMR Measurement Conductance Salt Measurement Chemical Shift Relaxation Measurement Lithium Free or SSIP Short-lived SSIP Ion Pairs (9 Salts) (LiCl) (3 Salts) Sodium SSIP Collisional IP Short-lived SSIP Ion Pairs (NaI) (9 Salts) (NaBr) (3 Salts) Potassium Collisional IP Ion Pairs (3 Salts) (3 Salts) Rubidium Higher Ion Aggregate Ion Pairs (RbI) (2 Salts) Cesium Collisional IP Ion Pairs (9 Salts) (2 Salts) Thallium Contact IP (3 Salts) Tetraalkyl- Ion Pairs Ammonium (7 Salts) SSIP: Solvent-separated Ion Pair. Free: Free Ion. 114 methods yield direct information on first coordination sphere effects. However, different kinds of ion-ion interactions have been proposed for many salt/NMF systems. This shows that even in this high donor, high dielectric constant solvent (NMF), and even at low salt concentration, there may be some kind of ion-association. These results also lead us to think about the phenomena and nomenclature of ion pairs because there are several ambiguous terms --- contact ion pair, solvent-separated ion pair, collisional ion pair, transient contact ion pair, weak contact ion pair, short-lived contact ion pair and short-lived solventé separated ion pair. A thermodynamic ion pair arises in dilute solutions when. oppositely charged ions approach one another more closely than average, approach sufficiently close to affect the total level of electrostatic interactions. As mentioned in section 2.3-2, this type of ion pair cannot exist in a high dielectric constant solvent at low concentration. Instead, ions collide with other ions within a very short range and collisional ion pairs could be formed. They can be detected by the NMR method only, because NMR parameters are sensitive to very short-range interactions. Thus, the difference between thermodynamic and collisional ion pairs is in the time scale of association (150). A collisional ion pair is formed if the collision between ions lasts for a 'minimum’ time which is long compared with the general rate of motion within the radial distribution around a 115 given ion, whereas a thermodynamic ion pair is formed if there is a very long association time. C. - O 8 ONE I - B W 1- 39.81113}; Cesium-133 chemical shifts were measured for eighteen mixed electrolyte systems including lithium, sodium and cesium salts (chloride, bromide, thiocyanate, perchlorate, tetraphenylborate and nitrate) and three iodide salts (tetrapentylammonium, sodium and potassium iodide) in 0.10 M CsI/NMF solutions. The results are given in Tables 51-61 and Figures 23-26 (Cesium-133 chemical shifts are given relative to that of 0.10 M CsI). Cesium-133 chemical shifts also change linearly with the concentration of the salts added to 0.10 M CsI/NMF solutions for all systems studied. CsSCN was also mixed with 0.24 M and 0.48 M C31, in addition to 0.10 M CsI. The results are given in Table 62 and Figure 27 (Cesium-133 chemical shifts are given relative to those of 0.24 M CsI and 0.48 M CsI) . Again, cesium-133 chemical shifts change linearly with the concentrations. Concentration shifts for all systems (including those for pure cesium salts) are summarized in Table 63. 2. 218.999.91.911 Covalency, electrostatic polarization and short range repulsive forces have been previously considered as possible sources of chemical shifts in the single salt 116 Table 51. Cesium-133 Chemical Shifts of NMF Solutions of Mixed CsCl/0.10 M CsI and LiC1/0.10 M CsI. Concentration Concentration of CsCl(M) 6(10.08ppm) of LiCl(M) 6(i0.08ppm) 0.0273 0.58 0.0401 0.81 0.0603 1.41 0.1050 2.18 0.0879 1.91 0.1239 2.55 0.1319 2.73 0.2029 4.36 0.1550 3.56 0.2300 4.86 0.2593 5.71 0.3055 6.41 0.3940 8.41 0.4647 10.03 Table 52. Cesium-133 Chemical Shifts of NMF Solutions of Mixed NaCl/O.10 M CsI and NaTPB/0.10 M CsI. Concentration Concentration of NaCl(M) 6(30.08ppm) of NaTPB(M) 5(10.08ppm) 0.0402 0.81 0.0237 -1.43 0.1095 2.30 0.0602 -3.67 0.2173 4.73 117 Table 53. Cesium-133 Chemical Shifts of NMF Solutions of Mixed LiTPB/0.10 M CsI and CsTPB/0.10 M CSI. Concentration Concentration of LiTPB(M) 6(10.08ppm) of CsTPB(M) 5(10.08ppm) 0.0175 -0.87 0.0045 -0.14 0.0242 -1.25 0.0157 -0.89 0.0382 -1.93 0.0284 -1.55 0.0478 -2.49 0.0336 -1.88 0.0563 -2.93 0.0457 -2.29 0.0583 -3.00 Table 54. Cesium-133 Chemical Shifts of NMF Solutions of Mixed LiClO4/0.10 M CsI and NaClO4/O.10 M CsI. Concentration Concentration of LiClO4(M) a(:0.08ppm) of NaClO4(M) 5(10.08ppm) 0.0202 -0.09 0.0580 -0.16 0.0494- -0.21 '0.1070 -0.37 0.0729 -0.29 0.1548 -0.52 0.1081 -0.46 0.1768 -0.57 0.1264 -0.53 0.2622 -0.88 0.1504 -0.63 0.3022 -0.99 0.1701 -0.71 0.3430 -1.09 0.1988 -0.83 . 118 Table 55. Cesium-133 Chemical Shifts of NMF Solutions of Mixed CsClO4/0.10 M CsI and CsNO3/0.1O M CsI. Concentration Concentration of CsClO4(M) 6(10.08ppm) of CsN03(M) 5(i0.08ppm) 0.0323 -0.106 0.0205 0.04 0.0575 -0.137 0.0416 0.07 0.0708 -0.189 0.0705 0.12 0.0990 -0.272 0.1270 -0.354 Table 56. Cesium-133 Chemical Shifts of NMF Solutions of Mixed LiN03/0.10 M CsI and NaNO3/0.10 M CsI. Concentration. Concentration of LiN03(M) 6(10.08ppm) of NaNO3(M) 6(10.08ppm) 0.0384 0.00 0.0235 0.02 0.0566. 0.00 0.0465 0.03 0.0899 0.00 0.0700 0.04 0.1175 -0.04 0.1141 0.08 0.1842 -0.04 0.1724 0.13 0.2633 -0.04 0.2188 0.17 0.3119 -0.04 0.3018 0.21 0.3322 -0.04 0.3618 0.25 119 Table 57. Cesium-133 Chemical Shifts of NMF Solutions of Mixed LiBr/0.10 M CsI and NaBr/0.10 M CsI. Concentration Concentration of LiBr(M) 6(:0.08ppm) of NaBr(M) 5(10.08ppm) 0.0115 0.31 0.0165 0.44 0.0691 1.57 0.0544 1.50 0.0835 2.18 0.0729 1.99 0.1186 3.11 0.0972 2.74 0.1307 3.49 0.1176 3.30 0.1480 3.92 0.1399 3.86 0.1624 4.30 0.1613 4.55 0.1773 4.80 0.1987 5.60 Table 58. Cesium-133 Chemical Shifts of NMF Solutions of Mixed CsBr/0.10 M CsI and LiSCN/0.10 M CsI. Concentration Concentration of CsBr(M) 5(10.08ppm) of LiSCN(M) a(¢0.08ppm) 0.0249 0.78 0.0308 0.24 0.0566 1.61 0.0484 0.37 0.0905 2.60 0.0723 0.57 0.1022 3.01 _0.0923 0.74 0.1262- 3.67 0.1584 1.27 0.1433 4.09 0.1707 1.36 0.1835 5.33 0.1992 1.56 0.2129 6.24 0.2407 1.89 120 Table 59. Cesium-133 Chemical Shifts of NMF Solutions of Mixed NaSCN/0.10 M CsI and CSSCN/0.10 M CsI. Concentration Concentration of NaSCN(M) 5(.~I_-0.08ppm) of CsSCN(M) 6 (10.08ppm) 0.0204 0.20 0.0236 0.24 0.0537 0.51 0.0416 0.41 0.1326 1.23 0.0730 0.74 0.3071 2.88 0.1068 1.07 0.4040 3.76 0.1343 1.40 0.5212 4.95 0.1545 1.57 0.6149 5.73 0.1759 1.70 0.2262 2.31 Table 60. Cesium-133 Chemical Shifts of NMF Solutions of Mixed NaI/0.10 M C31 and KI/0.10 M CsI. Concentration Concentration of NaI(M) I5(¢p.08ppm) of KI(M) 6(10.08ppm) 0.0240 0.88 0.0199 0.67 0.0717 2.45 0.0506 1.83 0.0927 3.28 0.0672 2.32 0.1348 4.60 0.0979 3.07 0.1541‘ 5.35 0.1657 5.55 0.1721 6.01 0.1958 6.70 0.1968 6.92 121 Table 61. Cesium-133 Chemical Shifts of NMF Solutions of Mixed TPAI/0.10 M CsI. Concentration of TPAI(M) 5 ($0.08ppm) 0.0146 0.49 0.0403 1.42 0.0728 2.39 0.0988 3.39 Table 62. Cesium-133 Chemical Shifts of NMF Solutions of Mixed CsSCN/0.2434 M CsI* and CsSCN/0.4816 M ** CsI Concentration Concentration of CsSCN(M)* 5(10.08ppm) of CsSCN(M)** I (10.08ppm) 0.0139 0.16 0.0319 0.37 0.0387 0.33 0.0568 0.62 0.0780 0.83 0.1063 1.11 0.0982. 1.07 0.1503 1.53 0.1126 1.16 0.2309 2.44 0.1340 1.40 0.3136 3.26 0.1919 2.06 0.4744 5.00 0.3814 3.97 0.7422 7.81 122 O l O‘) l l\) l .// L Cesium—133 Chemical Shift(ppm) '4 A A L A A A A J Gr u ‘- fl w f ‘3 x LiBr _2 d 0 LiCl x LiSCN . . A LJNO3 a LKNO4 o lJTPB "6 T7 I I I I C10 0.1 0.2 0.3 004 0.5 Concentration of Lithium sdlts Added to 0.1 M Csl Solution in NMF Figure 23. Plot of Cesium-133 Chemical Shifts vs. Concentrations of Lithium Salts Added to 0.10 M CsI Solutions in NMF. 123 NaBr NaCl NaSCN NaNO3 NaClO4 NaTPB I I I ”*1 0.00 0.15 0.30 0.45 0.60 Cesium—133 Chemical Shift(ppm) OUDXOBK ‘—6 Concentration of Sodium Salts Added to 0.1 M Csl Solution in NMF Figure 24. Plot of Cesium-133 Chemical Shifts vs. Concentrations of Sodium Salts Added to 0.10 M CsI Solutions in NMF. 124 8— A 6‘ E Q. O. E“ 4- 33 0 (fl .8 2- E Q) .C O 0‘0 t—r—fi 1“) WM N7 l E —2— .3 X CsBr (3 0 CsCl _4_ X CsSCN A CsNO3 Cl CsC|O4 -6 O CsTPB I I l I I j 0.0 0.1 0.2 0.3 Concentration of Cesium Salts Added to 0.1 M Csl Solution in NMF Figure 25. Plot of Cesium-133 Chemical Shifts vs. Concentrations of Cesium Salts Added to 0.10 M CsI Solutions in NMF. 125 8—~ E o. e- 6— {E m .. B .9 t 4— _c U 1") Pf) -4 | E 2— '5 m U 1 A NaI/0.1MCsl a Kl/O.1MCsl o TPAI/0.1MCsl 0 I I I I 0.00 0.05 0.10 0.15 0.20 Concentration of Iodide Salts Added to 0.1 M Csl Solution in NMF Figure 26. Plot of Cesium-133 Chemical Shifts vs. Concentrations of Iodide Salts Added to 0.10 M CsI Solutions in NMF. 126 8-— E o. e- 6- {E (f) 1.11 E .2 GE. 4_ .c U “3 m -1 l E 2— '5 m U " A CsSCN/0.48MCSI a CsSCN/0.24MCSI 0 o CsSCN/0.1OMCsl I I I *7 0.0 0.2 0.4 0.6 0.8 ‘ "Concentration of Thiocyanate Salts Added to Csl Solution in NMF Figure 27. Plot of Cesium-133 Chemical Shifts vs. Concentrations of CsSCN Added to 0.10 M, 0.24 M and 0.48 M CsI Solutions in NMF. 127 Table 63. Parameters of Linear Regression for Mixed Salt Systems in NMF. Concentration Correlation System Shift(ppm/M) Coefficient CsCl 19.83 0.9996 CSC1+0.10MCSI 22.05 0.9982 NaC1+0.10MCSI 21.95 0.9961 LiCl+0.10MCSI 21.62 0.9998 CsBr 30.29 0.9997 CsBr+0.10MCsI 29.05 0.9997 NaBr+0.10MCsI 28.37 0.9998 LiBr+0.10MCsI 27.36 0.9979 CsTPB -59.50 ~0.9998 CSTPB+0.10MCSI -52.79 -0.9930 NaTPB+0.lOMCsI ” -62.01 -0.9999 LiTPB+0.10MCSI -52.47 -0.9997 CSClO4 -1.34 -0.9840 CSC104+0.10MCSI -2.73 -0.9904 NaClO4+0.10MCsI -3.26 -0.9979 LiC104+0.10MCsI -4.18 -0.9997 128 Table 63 (Cont’d). System Concentration Correlation Shift(ppm/M) Coefficient CSNO3 2.67 0.9971 CSNO3+0.10MCSI 1.67 0.9984 NaNO3+0.10MCSI 0.70 0.9969 LiN03+0.10MCSI 0.00 ------ CsSCN 10.38 0.9994 CSSCN+0.10MCSI 10.09 0.9987 NaSCN+0.10MCSI 9.37 0.9999 LiSCN+0.10MCsI 7.91 0.9998 CSSCN+0.24MCSI 10.47 0.9994 CSSCN+0.48MCSI 10.51 0.9999 CsI 36.15 0.9995 NaI+0.10MCsI 34.83 0.9996 KI+0.10MCSI 33.72 0.9984 TPAI+0.10MCSI 33.87 0.9991 129 solutions, and only the last enables a satisfactory interpretation. of the experimental data. In *the present case it is easy to see that covalency and electrostatic effects are not plausible mechanisms. The short range repulsive forces present during direct collisions between the ions are again suggested as the most likely mechanism. The molar molecular shift for the salts added to 0.10 M CsI can be defined as the concentration shift difference between the mixed electrolyte system and 0.10 M CsI/NMF (the concentration shift of 0.10 M CsI/NMF is assumed to be zero). Since cesium iodide has the largest concentration shift (or the strongest overlap between cation and anion), molar molecular shifts for all the systems studied are all negative (Table 64). The values clearly show that the order of increasing shielding shift molar (or' decreasing molecular shift) is in agreement with that in the case of Table 64. Molar Molecular Shift (ppm/M) for Mixed Electrolyte Systems in NMF. 1' Cl' Br" TPB- c104' N03- SCN- Cs+ 0 -14.10 -7.10 -88.94 -38.88 -34.48 -26.06 Na+ -1.32 -14.20 -7.78 -98.16 -39.41 -35.45 ~26.78 Li+ ---- -14.50 -8.79 -88.62 -40.33 -36.15 -28.24 130 single cesium salt solutions: I" < 8:" < 01‘ < SCN- < N03“ < 0104' < T98". The molar molecular shifts are also slightly cation dependent. Except for the case of TPB-, all systems show shielding shift increases in the order: Cs+ < Na+ < Li+. The dependence of the cation chemical shifts on the other types of cation in NMF solution indicates that the effects are quite specific. The existence of interactions between ions of like charge has previously been suggested by thermodynamic, calorimetric (151—155) and NMR measurements (156). If the collision frequency remains essentially constant, ‘which. is likely for interactions between oppositely charged ions, then replacement of some cesium cations by sodium or lithium ions will lead to a smaller degree of mutual overlap and a upfield shift of the cesium- 133 resonance. This type of process will involve interactions between ions of like charge. For the CsSCN/Csl system, the Cs-133 chemical shift was measured as CsSCN was added. to different concentrations of cesium. iodide. The results .show that the concentration shift is essentially constant. The processes determining the chemical shifts are more or less independent of the total concentration. Tetraphenylborate salts are always interesting salts. In single electrolyte systems, the cesium-133 resonance of cesium tetraphenylborate always shows a very large upfield shift no matter what the solvent is. The magnitude of the 131 shifts (absolute value) are always as large as those of cesium iodide (downfield shift) . However, sodium tetraphenylborate shows no concentration dependence of the sodium-23 chemical shift in a variety of solvents eventhough sodium iodide still has the largest shift. This implies that for sodium tetraphenylborate the anion doesn't penetrate the cation solvation sphere and no significant chemical shift occurs. On the other hand, in the NaTPB/CsI mixed electrolyte system there is an unusually large molar molecular shift (absolute value) compared to those in the CsTPB/CsI and LiTPB/CsI systems. Again, this might be explained on the basis of the phenomena of cation-cation interactions. Since sodium ions don't interact with tetraphenylborate ions, then cesium ions have more chances to interact with sodium ions relatively. This will lead to a more negative value of the molar molecular shift relative to that of cesium iodide. D. - ON N 8 ION 1. W 1-1 1W 1.1-1 WW Solutions of LiSCN (0.06 to 0.60 M concentration range), NaSCN (0.05-0.94 M), CSSCN(0.05-0.92 M) and TIAASCN (0.06-0.50 M) in NMF were studied in the mid-ir region. The Peak position of the C54 stretching vibration for all of 1 the salts studied was 2057 i 1 cm- (Table 65), ie., the Peak frequency of the v mode of SCN- in NMF is independent 1 132 Table 65. Peak Positions v (cm-1) and Half-height Width (Kl/2) (cm-1) of CEN Stretching Bands of Four Thiocyanate Salts in NMF. Salt Concentration 11(cm-1) H1/2(cm-l) LiSCN 0.0560 2057.5 30 0.0823 2057.5 33 0.1946 2057.5 33 0.2376 2057.5 34 0.3162 2058.1 33 0.4263 2058.1 33 0.5775 2057.5 32 0.6040 2057.5 33 0.7951 2058.1 33 NaSCN 0.1034 2056.8 31 0.2122 2056.8 32 0.2815 2057.5 32 0.3989 2057.5 31 0.4441 2057.5 32 0.8024 2058.1 33 0.9412 2057.5 32 CsSCN 0.0527 2056.8 30 0.0947 2057.5 32 0.2181 2056.8 31 0.2820 2056.8 33 0.4011 2056.8 32 0.4916 2056.8 31 0.9239 2056.8 31 TIAASCN 0.0545 2056.8 31 0.1639 2056.8 32 0.2587 2057.5 32 0.4632 2057.5 32 133 of the nature of the cation and the solute concentration. This suggests the absence of contact ion pairs in these solutions: the thiocyanate ions are highly solvated. The GEN band half-height width for all of the solutions investigated is 32 i 2 cm-l. From this very broad band it can be concluded that there is probably a nonuniform environment surrounding the thiocyanate anions. Therefore mathematical deconvolution was used. As shown in Figure 28, three bands appear for all systems studied following the deconvolution. Their positions are 2045 i l cm-l, 2055 i 1 cm'1 and 2066 i 1 cm?1 . Since the positions of these three components are cation independent, the multiplet structure must result from interactions between the thiocyanate anions and solvent molecules. Excess crown ethers 12C4, 18C6 and cryptand C222 were added to LiSCN (0.06 M and 0.43 M), NaSCN (0.05 M and 0.44 M) and CsSCN (0.05 M and 0.40 M) solutions respectively. These macrocyclic molecules are known to efficiently complex the particular cations. Deconvolution shows that addition of the complexing reagents produces no effect on the systems studied: the positions of the three bands are unchanged. This shows again that on the infrared timescale there is essentially no cation-anion interaction in alkali thiocyanate NMF solutions. FTIR spectra of 0.24 M LiSCN, 0.27 M NaSCN and 0.40 M CsSCN solutions were also obtained at different temperatures from 264K to 328K. The deconvolution results are shown in Figure 29. It Can be 134 ‘ h 1 F T I 1950 2000 2050 2100 2150 Wavenumber(cm—l) Figure 28. Deconvolution of the C N Stretching Band of Thiocyanate Salts in NMF. 135 328K 298K 266K 17 I I I . I ‘7 .1950 2000 2050 2100 2150 Wavenumber(cm—l) Figure 29. Deconvolution of GEN Stretching Bands of Lithium Thiocyanate (0.24 M) in NMF at Three Temperatures. 136 clearly seen that the 2045 cm-1. peak changes its shape with changing temperatures whereas there is no change in shape or position of the other two components. The results confirm that several species exist in the systems studied. 1.1-2 Compute; Resolution of Thiocyanate Bang; The infrared contours given by the GEN stretching vibration of the thiocyanate ion in NMF do not have features that clearly indicate the individual components of these composite bands. Resolution of these contours was based on fitting their variations in spectra of solutions having different compositions. Each contour was fitted by a systematic series of approximations beginning with manual adjustment of band parameters until contours in spectra of several different solutions were matched by changing only the height of the components and ending with computer refinement (48). The quality of a match was visually judged by a display of the difference between the experimental and synthetic contours. The band shapes were obtained from spectra of TIAASCN solutions that contain primarily the unassociated SCN- ions. The EN bands were fit by synthetic bands that had 358 Lorentzian and 65% Gaussian character. After an iterative fitting' procedure for spectra. of several solutions to determine the positions and widths, the heights of the three bands could be varied independently to fit the contours given by several other concentrations. The final envelope fitting was again done by the computer with 137 the positions of the component bands fixed (2045.0 cm-l, 1 1). Curve fitting results are given in Tables 66-67 and 2054.5 cm” and 2065.5 cm- Figure 30. The results show that the three-peak fit is better than a two-peak fit, and that the half-height widths for the three peaks are relatively constant (30 i; 2 cm-l, 1 and 12 i 1 cm-1 respectively). Plots of 15 i, 1 cm- relative integrated intensities of these three peaks versus thiocyanate concentrations and solution temperature are given in Figures 31-36. It seems that for lithium, sodium and cesium thiocyanate there is no temperature dependence of the relative intensity (-9 - 55°C). This shows the uniqueness of solvation structures of thiocyanate anions in NMF. Kabisch and Klose (157), from their studies of the temperature dependence of the N-H band in NMF, also found that NMF liquid structures are almost temperature independent over the temperature range 295-390K. However, all three thiocyanate salts do show some concentration dependence of the relatiVe intensities of the three peaks. With increasing salt concentration, the intensity of the 1 1 2054.5 cm- peak increases whereas that of the 2065.5 cm- 1 peak decreases. The intensity of the 2045.0 cm- peak is essentially concentration independent. It is difficult to provide assignments for these three peaks. However, the 2054.5 cm-1 peak is the dominant species in the system. Thus, they must represent free thiocyanate anions. Thiocyanate ions may be wrapped around 138 Table 66. Curve Fitting Results of Thiocyanate Salts in NMF. Salt Concentration Relative Intensity(%) Chi- (M) lst’ 2nd* 3rd* Square LiSCN 0.0560 11.41 59.03 29.56 0.000298 0.0823 10.67 61.61 27.72 0.001984 0.1946 5.93 74.26 19.81 0.000745 0.2376 5.23 72.49 22.28 0.000480 0.3162 1.93 73.47 24.60 0.004981 0.4263 6.76 64.59 28.65 0.000636 0.5775 8.63 53.90 37.47 0.001094 0.6040 11.35 51.90 36.75 0.000773 0.7951 11.39 46.95 41.66 0.003362“ NaSCN 0.1034 13.10 70.37 16.53 0.000028 0.2122 12.22 69.97 17.81 0.000148 0.2815 10.05 69.66 ' 20.29 0.000345 0.3989 10.97 67.55 21.48 0.001410 0.4441 10.94 67.18 21.88 0.002871 0.8024 10.39 65.53 24.08 0.002050 0.9412 10.04 62.59 27.37 0.007214 CsSCN 0.0527 7.62 77.86 14.52 0.000788 0.0947 9.94 73.71 16.35 0.000293 0.2181 12.34 68.87 18.79 0.003519 0.2820 16.51 65.04 18.45 0.000509 0.4011 16.05 63.60 20.35 0.002272 0.4916 14.47 66.94 18.59 0.000978 0.9239 11.84 69.33 18.83 0.002371 * lst --- 2045.0 cm"1 2nd --- 2054.5 cm-: 3rd --- 2065.5 cm- 139 Table 67. Curve Fitting Results of Thiocyanate salts in MMF at Different Temperatures. Solution Temperature Relative Intensity (%) Chi- (OC) lst* 2nd* 3rd* Square 0.2376 M -9.0 6.58 70.65 22.77 0.001977 LiSCN -1.0 7.14 70.18 22.68 0.001804 5.0 7.33 70.09 22.58 0.001439 13.0 11.66 66.68 21.66 0.004171 25.0 5.23 72.49 22.28 0.000480 34.0 9.99 66.59 23.42 0.003517 41.0 9.07 66.37 24.56 0.001080 48.0 9.08 65.84 25.08 0.002122 55.0 8.72 67.55 23.73 0.002167= 0.4885 M -1.0 11.61 66.81 21.58 0.001381 NaSCN 10.0 9.51 67.94 22.55 0.006390 15.0 12.68 66.77 20.55 0.001581 25.0 10.04 69.31 20.65 0.001292 30.0 12.03 67.76 20.21 0.001247 39.0 8.13 70.54 21.33 0.007727 44.0 11.08 69.22 19.70 0.003326 0.4011 M -7.0 14.30 69.69 16.01 0.002437 CsSCN 0.0 12.15 70.80 17.05 0.002144 5.0 10.75 72.22 17.03 0.002052 15.0 12.10 71.03 16.87 0.001871 25.0 16.05 63.60 20.35 0.002272 38.0 16.30 66.41 17.29 0.002331 45.0 10.09 65.78 24.13 0.001855 '* lst --- 2045.0 cm'1 2nd --- 2054.5 cm'1 3rd --- 2065.5 cm'1 140 (18— C16— 8 2054.5 C o {3 (14— o 3 2065.5 < 0.2- \ 2045.0 0.0 ' l b '2 1 r l ' l 1950 2000 2050 2100 2150 Wavenumber(cm - 1) Figure 30. Resolved Curves of the CEN Stretching Band for Thiocyanate Salts in NMF. 141 100-1 0 2054.5cm—1 a 2065.5cm—1 A 2045.0cm—1 802 g: 1 ,3 604 .° m C o B 0 .E o G) a s 40— . 2 O m a 7 00 o 20— 11° 0 A I l l l 0.0 072 0.4 0.6 0.8 1.0 Concentration of LiSCN (M) Figure 31. Plot of Relative Intensities of Three (24 Resolved Bands vs. Concentrations of LiSCN in NMF. 142 100-— O 2054.5cm—1 D 2065.5cm—1 A 2045.0cm—1 80—~ g d o 0 ° 00 :1? 60- . m c .93 .E m .5 40— .2 w 0: Cl 20‘ 0 0° 0 l l 0.0 0.2 0.4 06 0:8 T0 Concentration of NaSCN (M) Figure 32. Plot of Relative Intensities of Three C'=‘N Resolved Bands vs. Concentrations of NaSCN in NMF. 143 100-1 o 2054.5cm—1 . CI 2065.5cm-1 A 2045.0cm—1 80- o 4:" 60— m c 8 £2 . m L; 40- .2 m m a 20‘ 0 o a a c: O I l l l l 0.0 0.2 0.4 0.6 0.8 1.0 Concentration of CsSCN (M) Figure 33. Plot of Relative Intensities of Three GEN Resolved Bands vs. Concentrations of CsSCN in NMF. 144 100.. O 2054.5cm—1 D 2065.5cm—1 A 2045.0cm-1 80—< E o o o o o o o .43? 60— (f) C .3.) .E m :3 40- .2 m 0: 20— ° ° ° 0 l I I i I i 1 —10 10 30 50 Temperature (°C) [0.2376 M uscn] Figure 34. Plot of Relative Intensities of Three can Resolved Bands vs. Temperatures in 0.24 M LiSCN/NMF Solution. 145 100-— O 2054.5cm—1 Cl 2065.5cm—1 A 2045.0cm—1 80~ g 4 ° 0 o o o 0 ° ,3? 60— m C 2 £5 . m .3 40— 2 0 Q: . 20‘ o o o a 0 D o O I I f I I I fi -10 10 30 50 Temperature (°C) [04885 M NaSCN] Figure 35. Plot of Relative Intensities of Three CEN Resolved Bands vs. Temperatures in 0.49 M NaSCN/NMF Solution. 146 100— O 2054.5cm—1 D 2065.5cm—1 A 2045.0cm—1 80-< g 0 O O .4? 60~ U) c B m .3 4o— .9 m 0: . 20“ o 2 o 0 o A 2 O I I ' I ' I fi -10 1O 30 50 Temperature (°C) [0.4011 M CsSCN] Figure 36. Plot of Relative Intensities of Three GEN Resolved Bands vs. Temperatures in 0.40 M CsSCN/NMF Solution. 147 by the flexible NMF chain structure or thiocyanate ions may be solvated through hydrogen bonds between either end of the thiocyanate ion (158) and the terminal N—H group of the NMF molecules, with cations attached to the oxygen (and nitrogen) atoms of the NMF molecules. Because of their low 1 intensities, the 2045 cm- and 2065.5 cm'1 peaks could represent the less solvated thiocyanate ions. For the 2045 cm.1 peak its relatively low wavenumber position may reflect less triple bond character, and the possibility that the nitrogen atom of the thiocyanate anion act as a bidentate or tridentate ligand. 1-1-3 WWW: mm Some studies of the solvent vibrations were also carried out. The absorption bands due to the stretching vibrations of the N-H, c=o and C-NI groups of NMF in 1% NMF/acetonitrile under the influence of dissolved LiSCN (the concentration of the saturated solution is less than 0.5 M), NaSCN, CsSCN(sparingly soluble) and Iron(III) perchlorate were examined. Acetonitrile was used as a medium in order to break the hydrogen bonds among NMF molecules since this hinders observation of the effects being studied. The results are shown in Table 68. From the observed shifts it is clear that the hydrogen bonds between NMF molecules are broken in the mixed solvent solution. For example, the N-H band shifts to higher wavenumber (+101 148 Table 68. Wavenumbers (cm-1) of N-H, c=o and C-N Stretching Bands in 1% (volume) NMF/MeCN Mixtures. NMF Absorption Bands N-H c=o C-N Solution y Au v Av v Av NMF 3301 --- 1667 --- 1243 --- 1% NMF/MeCN 3402 --- 1690 --- 1223 --- 0.15M LiSCN 3336 -16 1637 -3 1245 +17 0.38M LiSCN 3331 -21 1630 -10 1247 +19 sat'd LiSCN 3379 -23 1679 -11 1249 +21 0.10M NaSCN 3399 -3 1690 0 1229 +1 0.22M NaSCN 3397 -5 1639 -1 1231 +3 0.37M NaSCN 3391 -11 1636 -4 1237 +9 0.52M NaSCN 3337 -15 1635 -5 1240 +12 0.98M NaSCN 3379 -23 1633 -7 1242 +14 sat'd CsSCN 3403 +1 1690 0 1229 +1 0.09H Fe(ClO4)3 ----* --- 1643 -42 1274 +46 * can’t be determined because of low signal/noise ratio. 149 cm"1 ) in 1% NMF/MeCN compared to that of pure NMF liquid. In this kind of ternary system (NMF, MeCN and salt), all three components can interact with one another. Perelygin (159) and Pominov e_t a]; (160) investigated the effect of salts on the amide absorption bands in the ternary system by comparing the absorption spectra of the binary system (amide and acetonitrile) . They found that the interaction of acetonitrile with the salts has no effect on the positions of the absorption bands of amide by interaction with salts. The positions of the three bands of NMF change with salt concentration. The more salt added, the greater is the shift. Comparisons of the three stretching bands among Perelygin’s (123), Pominov's (161) and this work for different lithium and sodium salts are given in Table 69. The wavenumber shifts for the C-N and c=o stretching bands are the results of the coordination with the metal ions, mainly through the oxygen atoms and possibly the nitrogen atoms (123,159) . The wavenumber shifts of both bands are dependent on the charge densities of metal ions. The largest shifts of both bands are found in the case of li‘e(ClO4)3 which has the highest charge density of metal ions studied. Similarly, lithium ions have higher charge densities than sodium ions, so that the larger wavenumber shifts are found in the cases of lithium salts. The N-H stretching bands are very broad (linewidth z 50 cm-l). With increase in the salt concentration, this band Table 69. Wavenumber Shifts Stretching Bands of NMF in Mixtures. Solution a NaClO4 NaSCN NaI a Mg(ClO4)2 Fe(Clo4)3 150 (cm-1 ) 0f N-H, C=0 and C-N 1% (volume) NMF/MeCN +21 +21 +25 +10 +14 +11 +35 +46 a: Perelygin and co-workers (reference 123). b: Pominov and Pavlova (reference 159). 151 moves toward lower frequencies, which results from the interaction between metal ions and the nitrogen atoms of NMF, only as indicated by Perelygin. This kind of interaction would be stronger for lithium ions than for sodium ions so larger shifts are observed for the lithium salts. In Table 68, lithium thiocyanate has the higher wavenumber shifts than sodium thiocyanate if their concentrations are equal. However, the wavenumber shifts are also dependent on the anions (Table 68). This indicates that different anions interact with the N-H group of NMF molecules to a different extent. It is hard to compare the magnitude of the wavenumber shift in different anion systems because the N-H band is very sensitive to the salt concentration, whereas the concentrations are different for the salts listed in Table 69. Bonner and Jordan (107) studied the N—H stretch of NMF for several salts with the same concentration (4 M) and found from the wavenumber shifts that the anions interact in decreasing order (Table 4): F'.’ > Cl'z Br- > 1" > N03- > c104" > PF6-. Halide salts can interact with the N-H proton via hydrogen bonding. Accordingly, thiocyanate ions could also form hydrogen bonds with the N-H proton of NMF. However, the hydrogen bonds must be weaker in thiocyanate systems than that of the NaI system. 152 1-2 EMBL§25§EISQQBS§ 1-2‘1 §:1_E!B_!23§B££!2n£§ Proton chemical shifts were measured for 2.11 M LiCl/NMF, 1.21 M: NaSCN/NMF and. pure NMF solutions. The results, given in Table 70, show that the chemical shifts of the methyl and aldehyde protons of NMF in the IJCl and NaSCN solutions shift to higher field compared to those of pure INMF. These shifts reflect ‘the interactions between metal ions and the oxygen and nitrogen atoms of the NMF molecules. The interesting signal is that of the N-H group. The N-H proton shifts downfield compared to that of pure NMF. It is known that protons involved in hydrogen bonding undergo resonance at a lower applied magnetic field than non-associated protons (162). Then it can be concluded that the order of the hydrogen bond strength in NMF decreases by: chloride ion > thiocyanate ion > NMF molecule. This is consistent with the results from 1% NMF b Table 70. Proton Chemical Shifts of NMF and Its Solutions. -c§3 -cgo -Ng Solution 6 A6 6 A6 NMF ‘2.75 ---- 3.09 ---- 7.91 ---- 1.21M NaSCN 2.67 -0.03 3.01 -0.03 7.93 +0.02 2.11M LiCl 2.61 -0.14 7.97 -0.12 8.19 +0.28 153 /MeCN. Additions of LiCl and NaSCN partially destroy the hydrogen bonds already existing in solution. In turn, the loss of structure will be compensated by a greater degree of ionic solvation. Further, in the LiCl and NaSCN solutions the linewidth of the N-H proton signal decreases and becomes more observable. This may result from the decrease in proton exchange or reduction in quadrupole effects. 1-2‘2 H:111_E:1é_QBQ.QL:1§.EBB.§2§§Q££E§E&£ Values of the N-l4, N-15 and Cl-35 chemical shifts of thiocyanate, nitrate and perchlorate salts at various salt concentrations, are given in ‘Tables 71-73. There is no concentration dependence of chemical shift for these three nuclei. The N-14 and Cl-35 chemical shift measurements are not sensitive probes of the ionic solvation because of their broad linewidths. Moreover, for nitrate and perchlorate salts, the nitrogen and chlorine atoms are surrounded by three and four oxygen atoms respectively. This makes the measurements even less sensitive because the nitrogen and chlorine atoms would not be easily disturbed, which leads to very small chemical shift changes. The N-15 chemical shift measurement should be a better probe because of its very narrow linewidth. However, the N-15 chemical shifts of LiSCN-15 are still concentration independent because lithium ions are well solvated and the interaction between thiocyanate ions and the N-H proton of NMF is very . weak in this concentration range (0.04-0.34 M). 154 Table 71. N-14 Chemical Shifts of Thiocyanate and Nitrate Salts in NMF. N-14 Chemical Shift (il.8ppm) Salt Concentration(M) -SCfl C-H(NMF) -303 LiSCN 0.1102 -166.6 -263.7 ----- 0.4231 -165.1 -262.3 ----- 0.6437 -166.6 -262.7 ----- NaSCN 0.1135 -166.3 -267.0 ----- 0.2356 -167.2 -266.9 ----- 0.5373 -166.3 -266.6 ----- 0.3051 -166.3 -266.5 ----- CsSCN 0.1032 -165.9 -266.7 ----- 0.2639 -164.5 -265.4 ----- 0.4095 -164.1 -265.3 ----- 0.7652 -163.6 -265.3 ----- TIAASCN 0.0325 -164.5 -266.6 -311.5* 0.1127 -163.6 -266.5 -311.5* 0.2423 —165.1 -266.5 -311.5* NaNO3 0.0302 ------ -265.5 3.0 0.2233 ------ -265.7 3.0 0.4672 ------ -265.0 3.0 0.7073 ------ -265.0 2.9 CsN03 0.0433 ------ -266.5 3.0 0.0929 ------ -266.5 3.0 * -n(isoamyl)4+ 155 Table 72. N-15 Chemical Shifts of LiSCN-15 in NMF. Concentration(M) N-15 Chemical Shift(i0.04ppm) 0.0423 -162.60 0.1664 -162.58 002146 -162058 0.3387 -162.56 Table 73. Cl-35 Chemical Shifts of Perchlorate Salts in NMF. Salt Concentration(M) 6(io.14ppm) LiClO4 0.1870 -125.84 0.1104 -125.81 0.0630 -125.81 NaClO4 0.2899 -125.81 0.2287 ~125.81 0.0939 -125.81 CSCIO4 0.1835 -125.67 0.0919 -125.67 0.0695 -125.67 TlClO4 0.5023 -126.08 0.3594 -125.95 0.2067 -125.81 0.1302 -125.81 0.0648 -125.67 156 2- 9959.13.91.99: The results presented above clearly show that solvation of cations in NMF is through ion-dipole interaction with =0 and probably C-N groups, while that of anions is through interactions with the N-H groups. Halide and thiocyanate ions form hydrogen bonds with N-H groups, whereas perchlorate and nitrate ions interact with protons in a different way. Infrared measurements of thiocyanate systems do not reveal ion-ion interactions, unlike the NMR measurements. In fact, the infrared time scale is much faster than that of NMR and ion-ion interactions should also be observable using the infrared technique. The possible answer for this is that the ion-ion interactions in NMF are very weak and the solvation structures of the thiocyanate ions are not very much perturbed: this might lead to a very small change in the infrared spectra, unobservable under the experimental conditions. E. 0N8 F 38 The studies already stimulate the following suggestions for further studies: 1) A systematic study on ion-solvent interactions by H- 1 and C-13 NMR measurements. 2) Halogen NMR measurements on halide salt solutions in NMF. This should provide more information about interactions between anions and cations or solvent molecules. 3) N-14 linewidth measurements have been used to study 157 the ion pair formation in the TlNo3/liquid ammonia system. It must be interesting to use the same technique to study the TlNOB/NMF system in which contact ion pairs are formed. 4) Infrared measurements of perchlorate and nitrate NMF solutions. 5) Since solvent-separated ion pair has been found in formamide system by infrared studies on the N-D band, it must be also interesting to carry them out in NMF system. 6) Raman studies of highly concentrated thiocyanate solutions in NMF. Hopefully, some bands resulting from cation-anion interactions could be found. CO 159 A- IEIBQDHQIIQ! Previous studies in our laboratories (71,72,78-84) and elsewhere (65,66,68-70,73,75,76) have shown that nuclear magnetic resonance of alkali nuclei offers a very sensitive technique for the studies of changes in the immediate chemical environment of the alkali ions in solution. The chemical shifts and linewidths of the resonances can give information about ion-ion, ion-solvent, and ion-ligand interactions. During the past decade alkali metal NMR has been used extensively to study the thermodynamics and kinetics of the complexation reaction between alkali metal ions and crown ethers and cryptands. Complexation between crown ethers or cryptands and metal ions has not been previously studied in N- methylformamide solutions. The work presented in this chapter describes the complexation reaction of cesium ions with a series of crown ethers and cryptands, such as 18C6, DBl8C6, DA18C6, DC18C6, D321C7, DBZ4C8, DBZ7C9, C222, 0221 and C211, in NMF by using cesium-133 and C-13 NMR measurements. These studies not only provide the stability constants of different complexes but also give useful information about the effects of the solvent, and the different donor atoms of the ligands. 160 B. REEQLQQ MD QISCUSBLOH 1. m e at o o e um Ions b 806 n 8 b tut na o 1.1 139; Complexation between 18C6 and two cesium salts (0.05 M cesium ‘tetraphenylborate and 0.05 M (and 0.01. M: cesium iodide ) were studied. The cesium-133 chemical shift was determined as a function of the ligand to cesium. mole ratio. The results are listed in Tables 74-76 and Figure 37. A downfield shift followed by a break and then an upfield shift which gradually approaches a limiting value can be explained by the formation of a strong 1:1 complex followed by the addition of a second ligand to form a 2:1 sandwich complex (168) . It is known that the paramagnetic shift is determined by the overlap of donor's lone-pair electrons with p or d orbital of the cation (143). So the upfield. shift of the 2:1 complex ‘probably results from large decrease in the individual overlap integrals in the 2:1 complex as the short range repulsions are relaxed. The complexation formation constants were determined by the procedure described in the experimental part and Appendix II, and the results are given in Tables 77 and 78. Three effects are discussed as follows: Concentration effect: Different curves were found for 0.01 M CsI/18C6 and 0.05 M CsI/18C6 systems. In the case of 0.05 M CsI a stronger, sharp break ‘was observed. The difference in the formation constants (1:1 and 2:1) are not 161 Table 74. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (18C6)/(CsI) in NMF. [CsI] = 0.0502 + 0.0002 M. Mole Ratio Chemical Shift (18C6)/(Cs+) (+0.04ppm) 0 1.16 0.01 2.22 0.12 1.56 0.43 1.41 0.25 2.25 0.63 -0.27 0.38 2.46 1.04 -0.95 0.57 3.42 1.54 -3.38 0.77 4.28 2.04 -5.56 0.82 7.04 2.16 -11.25 0.88 7.48 2.22 -12.29 0.95 9.01 2.35 -14.87 1.01 9.89 2.35 -16.63 1.04 11.50 2.29 -19.00 162 Table 75. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (18C6)/(CsI) in NMF. [CsI] = 0.0101 + 0.0002 M. Mole Ratio Chemical Shift (18C6)/(Cs+) (+0.04ppm) 0 1.22 -1.43 1.26 0.20 1.42 -0.31 1.46 0.41 1.73 -0.19 1.46 0.61 2.03 0.32 1.36 0.31 2.54 0.74 1.26 0.91 3.05 0.95 0.95 0.96 4.06 1.05 0.63 1.02 7.61 1.05 -1.33 1.07 3.49 1.15 -1.95 1.12 9.22 1.15 -2.05 163 Table 76. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (18C6)/(CsTPB) in NMF. [CsTPB] = 0.0506 + 0.0002 M. Mole Ratio Chemical Shift (1806)/(Cs+) (10.04ppm) 0 1.23 -4.31 0.35 0.10 1.29 -4.07 0.73 0.22 1.96 -3.33 -0.39 0.41 2.39 -2.14 -1.95 0.57 3.75 -1.03 -5.56 0.73 5.50 -0.21 -9.92 0.35 6.67 0.29 -12.60 0.90 7.39 0.43 -15.23 1.00 9.16 0.79 -17.35 1.03 10.49 0.79 -19.42 1.09 12.10 0.35 —21.69 1.12 0.92 164 10— O n l “a. Cesium—133 Chemical Shift(ppm) _.1Q_. _20_ a 0131MCsl A 0.05MCsl o 0.05MC TPB —30 I I I I I I ' ' I 1 I [S I fi 0 3 6 9 12 Male Ratia(18C6)/(Cesium Ian) Figure 37. Plot of Cesium-133 Chemical Shifts vs. (18C6)/(Cs+) Mole Ratio in NMF. Table 77. Logarithms 165 of Formation Constants of 1:1 and 2:1 (18C6/Cs+) Complexes of Cesium Salts with 18C6 in Various Solvents. Dielectric Solvent Constant Gutmann Donor No. logK(1:l) logK(2:l) MeCN 37.5 14.1 >5 0.57 PC 65.0 15.1 4.18 1.04 AC 20.7 17.0 >5.30 1.53 DMF 36.71 26.6 3.95 0.39 DMSO 46.68 29.8 3.04 0.0 Py 12.4 33.1 >4 1.04 NMF 182.4 25 0.01MCsI 2.74610.001 0.210.? 0.05MCSTPB 2.7:0.1 0.0510.03 0.05HCSI 2.9:0.1 -0.1110.03 Except for NMF, {CsTPB} = 0.01 M and all data are taken from reference 78. 166 Table 78. Limiting Chemical Shifts of Cs+, Cs+(13c5) and Cs+(18C6)2 in Various Solvents. Solvent Cs+ [Cs+(l8C6)] [Cs+(18C6)2] (ppm) (ppm) (ppm) pc -36.5 -3.1¢0.2 ' -44.5¢0.3 AC -35.3 -6.4 -47+9 Py -32.4 10.2io.2 -43.0:0.2 ouso 63.0 23.610.4 -49¢24 our -0.3 3.37:0.05 ~48-210-7 . Mean 24.1 14.310.2 -53¢7 NMF -1.34 0.01xcsx 2.6:0.3 -43i62 0.05MCsI 3.1+0.1 -75¢4 0.05MCsTPB 2.7+0.2 -65¢3 Except for NMF, [CsTPB] = 0.01 M and all data are taken from reference 78. 167 significant within experimental error. This result is reasonable because the equilibrium constant should be concentration independent. However, the limiting chemical shifts of the 1:1 complexes are slightly different (about 1 ppm). From Chapter III, we note that for pure cesium salts in NMF the magnitude of the shifts is proportional to the salt concentration: 0.05 M CsI has larger paramagnetic shift than 0.01 M CsI in NMF ‘because of more ion-ion interactions. In the same way, one would expect that the 1:1 Cs+-18C6 complex in 0.05 M CsI could show a larger paramagnetic shift than that in 0.01 M CsI because cesium ions are not completely inside the 18C6 cavity, which leaves them accessible to iodide ions. The cesium ions in the 1:1 complex in 0.05 M CsI system have more interactions with iodide ions than those in the 1:1 complex in 0.01 M C31. The chemical shifts for 2:1 complexes in both cases are almost equal. This is understandable because in Cs+(18C6)2 complexes the cesium ions are sandwiched between two 18C6 molecules and‘have no interaction with the iodide ions. Thus the chemical shift of the 2:1 complex is concentration independent. It is? noted that the error is much larger in the 0.01 M CsI case because the total chemical shift change is very small (2 ppm), which makes the fitting difficult. Since there is no concentration dependence of formation constants, 0.05 M cesium ion concentration were used for all systems studied because the results would be more precise. 168 Anion effect: The 0.05 M CsTPB/18C6 system has the same shape of the curve as that of CsI/18C6 at the same Cs+ concentration. This result indicates that 1:1 and 2:1 complexes are also formed 'in this system. The formation constant is the same as that in the 0.05 M CsI system. The 1:1 complex in 0.05 M CsTPB has less paramagnetic shift than the 1:1 complex in 0.05 M CsI. This is expected because the tetraphenylborate anion is an electron- withdrawer, which makes the interaction between Cs+ in 1:1 complexes and anions weaker. There are differences in the formation constants and chemical shifts for 2:1 complexes between the C81 and CsTPB systems. It seems possible that the 2:1 formation constant for CsTPB(18C6)2 is larger than that in the CsI system because tetraphenylborate ions are less capable than iodide ions to compete with 18C6 molecules for’ 1:1 Cs+(18C6) complexes. However, as mentioned earlier chemical shifts for 2:1 complexes should be concentration and anion independent: yet there is some variation. Tw0 problems should be considered. One is that the chemical shifts of free cesium ions are not constant, whereas they are assumed constant in the program used to determine the formation constants. The other is that there are differences in the chemical shift change rate before reaching the top of the curves between the two cases, which results from the differences in concentration shifts between pure CsI and CsTPB solutions in NMF. These two problems lead to the different observed chemical shifts for 169 the 2:1 complexes. Solvent effect: In Tables 77 and 78, data from previous studies (78) of the complexation between the cesium ion and 18C6 in six other solvents are reported. There is no simple relationship between the formation constants and either the dielectric constant or the donor number of the solvent. The complexation reaction is essentially an‘ ion-dipole interaction, and donor number should play a more important role than the dielectric constant. Generally, solvents with lower donor number should have higher formation constants. However, NMF has approximately the same donor ability as DMF, and the formation constant in NMF is about ten times smaller than that in DMF. This might imply that there is some kind of interaction between 18C6 and NMF molecules, just as formamide molecules can form 2:1 complexes with 18C6 (169) via hydrogen bonding. Thus the formation constant in NMF is smaller than that in DMF. 1.2 M The attachment of two benzo groups on 1806 to form D81806 was expected to result in weaker complexes. This aromatic group decreases the 0---0 distance, which causes a decrease in cavity size (170). A mole ratio study of D818C6 complexes with CsI in NMF was made. The results (Tables 79 and 82, and Figure 38) show that the Cs+-D318C6 complex must be very weak because there is not even an inflection point in the plot of chemical shift versus mole ratio. The upfield shift is expected because benzo groups make the 170 Table 79. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (D818C6)/(CsI) in NMF. [CsI] = 0.0493 + 0.0002 M. Mole Ratio Chemical Shift (D818C6)/(Cs+) (+0.04 ppm) 0 0.01 0.21 -0.63 0.44 —1.30 0.54 -2.30 0.72 -3.23 0.31 -3.79 0.87 -4.04 0.97 -4.73 1.03 -5.04 1.10 -5.41 1.30 -6.53 -7.84 171 electron density of the cavity smaller and the interaction between D818C6 and Cs+ is weaker than that between NMF molecules and cesium ions. Again, the formation constant in DMF is about ten times larger than that in NMF. Interactions between DBlBCG and NMF molecules are possible. 1.3 m It would be interesting to test the effect of a complexing agent which does not have aromaticity: therefore dicyclohexyl-18C6 (DC18C6) was used. This ligand has five isomers, of which only two have been isolated (termed isomers A and B) . A mixture of these two isomers was used in this study. The results (Tables 80 and 82, and Figure 38) show that the extent of the downfield shift is much greater than that in the case of 18C6. This phenomenon may be caused by the higher rigidity of the DC18C6 framework which results in a larger overlap of the lone pair electrons of oxygen with the cesium ion. However this rigidity also leads to a drop in the stability of Cs+- DC18C6 complex as compared with the Cs+-18C6 complex (164). Again (Table 82), the formation constant in DMF is much higher than that in NMF. 1.4 951306 The replacement of two oxygen atoms by two sets of N-H atoms on 18C6 to form diaza-18C6 (DA18C6) is expected to result in weaker complexes too (76,171-173), because the electronegativity decreases as oxygen atoms are replaced by N-H groups. The results (Tables 81 and 82, and Figure 38) 172 Table 80. Variation of Cesium-133 Chemical Shifts with the (DC1BC6)/(CsI) Mole Ratio in NMF. [C81] 5 0.0493 + 0.0002 M. Mole Ratio Chemical Shift (DC18C6)/(Cs+) (+0.04 ppm) 0 0.01 0.13 1.13 0.29 3.31 0.46 6.54 0.53 3.97 0.97 15.76 1.00 16.63 1.09 13.93 1.13 19.00 1.22 20.99 1.37 22.43 1.99 23.40 2.72 29.27 3.13 29.39 4.04 29.33 5.00 29.33 173 Table 81. Variation of Cesium-133 Chemical Shifts with the (DA18C6)/(CsI) Mole Ratio in NMF. [CsI] = 0.0493 + 0.0002 M. Mole Ratio Chemical Shift (DA18C6)/(Cs+) (+0.04 ppm) 0 0.01 0.14 0.51 0.25 0.94 0.43 1.69 0.54 2.12 0.34 3.49 0.94 3.37 0.96 3.99 1.00 3.99 (1.12 4.61 1.27 5.13 1.51 6.30 1.95 8.10 2.25 9.03 3.02 11.34 3.64 13.77 5.40 18.56 174 30-— E a. 3 20— {E U) __ 1 a .2 g 10— .C a U 0 N? pr) I E o 3 0-4 ’6 m L) A—A DB18C6 a DA18C6 G—e DC18C -10 I , I , j 0 2 4 6 Male Ratio (Ligand)/(Cesium Ian) Figure 38. Plot of Cesium-133 Chemical Shifts vs. (Ligand)/(Cs+) Mole Ratio in NMF. 175 Table 82. Logarithms of the Formation Constants of Cesium Complexes with Crown Ethers in Various Solvents. Solvent D818C6 DC18C6 DA18C6 PC 3 4 ...... MeCN 1.54 > 4 2.48 AC > 3 > 4 ...... DMF 1.48 3.45 ------ DMSO 1.34 2.20 ------ PY 3.35 (1:1) > 5 ...... 2.36 (2:1) Me0H 3.55 (1:1) 4.25 ------ 2.92 (2:1) “20 0083 ---- ...... NMF (R z 0) 1.6+0.1 0.05+0.04 Except for NMF, all data are taken from reference 65. 176 show that the formation constant is much smaller than that of the Cs+-18C6 complex. 2- W The cesium cation has a diameter of 3.383 , which is just the right size to fit conveniently into the cavity of ligands D321C7, DBZ4C8 and D827C9, which range from 3.4 to 4.3K . Consequently, these Cs+-ligand complexes should be very stable. The results (Tables 83-85 and Figure 39) show that the cesium-133 chemical shift is essentially unaffected by changes in the mole ratio of ligand to cesium ions in the case of D321C7, and shifts upfield as expected in the cases of D824C8 and D827C9. There. are two possible explanations for the surprising result in the case of D321C7. One is that the chemical shifts of the complexes are very close to those of free cesium ions. However, this is very unlikely because the environment of complexed Cs+ should differ markedly from that of the free cation. Indeed, the environment of the cesium ions should be similar for all three ligand complexes. There is no reason for Cs+-D821C7 to have a very different chemical shift from those of other two complexes. The second explanation is more likely. There is a possibility of reaction between D821C7 and NMF molecules. As mentioned before, NMF might interact with 18C6 and its substituted analogs. However, in general the interaction is not strong enough to prevent formation of Cs+-crown ether complexes. Perhaps D821C7 has just the right size for NMF 177 Table 83. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (DBZlC7)/(CsI) in NMF. [CsI] = 0.0493 + 0.0002 M. Mole Ratio Chemical Shift (DBZlC7)/(Cs+) (+0.04ppm) 0 0.99 0.00 0.21 0.17 1.04 0.04 0.24 0.44 1.16 0.09 0.26 0.53 1.23 0.12 0.23 0.66 1.31 0.15 0.30 0.33 2.37 0.17 0.33 0.94 0.19 178 Table 84. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (DBZ4C8)/(CSI) in NMF. [CsI] = 0.0493 + 0.0002 M. Mole Ratio Chemical Shift (DBZ4C8)/(Cs+) (+0.04ppm) 0 0.94 0.00 -15.37 0.22 1.00 -4.35 -16.12 0.39 1.06 -7.65 -16.99 0.64 1.12 -11.32 —17.37 0.75 1.23 -13.33 -13.11 0.91 -15.37 Table 85. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (D827C9)/(CsI) in NMF. [C311 8 0.0498 1 0.0002 M. Mole Ratio Chemical Shift (D327C9)/(Cs+) (+0.04ppm) 0 0.93 0.00 -16.43 0.09 0.99 -2.11 -17.24 0.27 1.03 -5.35 -17.30 0.51 1.15 -10.39 -13.36 0.70 1.13 -13.63 -13.61 0.36 -15.62 ' 179 (3 4» p 1 1’ -121 CeSUnT—133 Chenflcalfflflfi(ppafl $6 I -—O 03 l A 082107 a 0824C8 _20 o 0827C9 T r 1 0.0 0.5 ' 11.0 1.5 Nkfle RaUa (Ugands)/(Ceskwnians) Figure 39. Plot of Cesium-133 Chemical Shifts vs. (Ligand) /(Cs+) Mole Ratio in NMF. f 180 molecules to form very strong complexes. Then cesium ions cannot compete effectively with NMF molecules for D821C7, and almost no Cs+-0821C7 complex will be formed. Cesium ions do form complexes with DBZ4C8 and DBZ7C9. Their formation constants are about the same as those in DMF (Table 86), which is different from what is found in the cases of 1806 and its analogs. This could be explained if D824C8 and D827C9 are too big for NMF molecules to form complexes. For large crown ethers, which are capable of forming stable three-dimensional 'wrap around' complexes with smaller cations (70,79,174), it can be expected that the size of the cation will influence strongly the extent of the complexation reaction. If the cation is too large, the three-dimensional structures cannot be formed and only some of the oxygen atoms can bond to the cation; consequently a weaker complex results. On the other hand, if the ring size is much larger than the cation, the ligand can still form the ’wrap around' structure, but in this case the oxygen atoms of the ligand are in close proximity and the resulting repulsive force will weaken the complex. The formation constant and the limiting chemical shifts obtained in the study for the complexation of cesium ions by D824C8 support the above conclusions. The cesium ion is still too large to form a ’wrap around’ complex. Thus the cation remains exposed to the solvent molecules, which results in the solvent dependence of the cesium-133 181 Table 86. Logarithms of the Formation Constants of Cesium Complexes with Crown Ethers in Various Solvents. Solvent 032107 0324c3 032709 PC 3.30 3.25 3.64 MeCN 3.95 3.94 3.39 AC 3.93 3.71 4.24 DMF 2.34 2.10 2.20 DMSO 1.72 1.61 1.33 Py 4.27 4.00 4.15 Me0H 3.96 3.65 3.52 NMF (K:= 0) 2.24+0.09 2.4+0.3 Except for NMF, all data are taken from reference 65. 182 resonance for the complexed cesium ion (Table 87). an the other hand, in the case of D827C9 the cesium ion is more effectively insulated from the medium as a result of the formation of a three-dimensional ’wrap around’ complex; the limiting chemical shifts of the complexed cesium ion are very nearly independent of the solvent (82). 3. WWW Complexation equilibria between 0222, C221 and 0211 and two cesium salts (0.05 M cesium iodide and cesium tetraphenylborate) were studied. The results are given in Tables 88-92 and Figure 40. In all cases, complexation of the cation results in a paramagnetic shift of the cesium- 133 resonance. For the same ligand/Cs+ mole ratio the magnitude of the chemical shift is in the order C222 > C221 > C211. Cryptands containing three-dimensional cavities of suitable size can form both inclusive and exclusive stable complexes with the metal ions (71). In the studies of complexation between C222 and two cesium salts (iodide and tetraphenylborate), the results show that there is no anion dependence of the formation constants, and of the limiting chemical shifts of Cs+-C222. The limiting chemical shifts of Cs+-C222 for both salts (227 + 22 ppm and 215 i 17 ppm) are very close to that of the inclusive Cs+-C222 complex (245 :3; 5 ppm). This implies that most of the complexes in NMF are inclusive cryptate complexes, so no anion dependence of the limiting chemical shifts of cryptate 183 Table 87. Limiting Chemical Shifts of 1:1 Complexes of Cesium Ions with D821C7, D824C8 and DBZ7C9 in Various Solvents. Limiting Chemical shift (ppm) Solvent 0821C? DBZ4C8 DBZ7C9 MeCN 8.17 -14.86 '22.34 AC -7.98 -29.24 ---- Me0H -19.07 -36.47 ---- DMF -0.92 -24.93 -24.19 DMSO 0.96 8.32 -20.14 Py 6.21 -20.75 -21.98 NMF ---- -22.98 -23.09 Except for NMF, all data are taken from reference 82. 184 Table 88. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (C222)/(CsI) in NMF. [CsI] = 0.0503 1 0.0002 M. Mole Ratio Chemical Shift (€222)/(Cs+) (ppm) 0 1.00 -0.01i0.04 137.59:0.12 0.11 1.08 ll.80:0.21 143.6910.12 0.28 1.18 40.74:0.21 149.9910.12 0.50 1.22 72.99:0.21 152.89i0.12¢ 0.66 1.68 99.03i0.21 167.67i0.11 0.69 1.95 101.72i0.16 172.11i0.11 0.82 2.35 116.71i0.l6 174.9010.11 0.87 4.08 120.74i0.16 179.66+O.11 0.95 6.22 132.42i0.16 181.3110.11 185 Table 89. Variation of Cesium-133 Chemical Shifts with the Mole Ratio (C222)/(CsTPB) in NMF. [CsTPB] = 0.0506 + 0.0002 M. Mole Ratio Chemical Shift (c222)/ 5 ----------- 224.3 NMF 2.27+0.01 2.3+0.2 (K:= 0) 227+22 (CsTPB) 215+17 (CsI) Except for NMF, all data are taken from reference 80. 190 macrocycle has a more widely open cavity and may be better able to accept the cesium ion than is the case with the C222 ligand (80). C221 forms a more stable complex with cesium ion in DMF than that in NMF even though they have almost the same donor number. As was mentioned earlier, there must be some kind of interaction between the ligands and NMF solvent molecules. The C-13 NMR spectrum of C221 in NMF was measured (Figure 41). The difference in C-13 chemical shifts between C-4 and C-5 in C221 at room temperature is 0.82 ppm (Table 93) . It has been reported (177) if that a difference in C-13 chemical shifts between C-4 and C-5 of C221 (Table 93 Solvent A) is indicative of specific solvent-ligand interactions. A portion of the MMF molecule (N-H or CH3) might penetrate into the cavity of the ligand and form hydrogen bonds with the heteroatoms of 0221. C. W The stabilities of complexes between cesium ions and crown ethers and cryptands decrease in the order: 18C6 > D827C9 z D824C8 z C221 z C222 > DC18C6 > DA18C6. DBl8C6, DBZlC7'.and C211 do not form complexes with Cs+. The formation constants are usually small because of the high donicity of NMF. Solvent-ligand interaction seems to be a very important factor in determining the stability of the complex. D821C7 might have just the right geometrical environment for NMF molecules, which would account for the absence of Cs+/D821C7 complexation. In the cases of 18C6, 191 ”2 . 4’, ‘ t. Q ~ 0 /“\ / :\ /\ :22. "f V0 V/N x / o o . \/ \\/ Figure 41. Carbon-13 NMR Spectrum of C221 in NMF. 192 Table 93. C-4 and C-5 Separation in 13C Spectra of Cryptand C221 in Various Solvents. Solvent A t(°0) a 4-55(ppm) Solvent 3 t(°0) 5 4-,5(ppm) NMF 24 0.81 DMF -53 0.0 FA 34 2.32 THF -88 0.0 H20 30 2.00 1,3-Dioxolane -93 0.0 Nitromethane -20 1.78 Py -39 0.0 Nitroethane —70 1.49 AC -90 0.0 Nitropropane -80 1.26 Toluene —80 0.0 MeCN -42 1.64 Anisole -37 0.0 MeCN 24 0.47 Methyl Acetate 34 0.0 MeOH -80 1.70 MeOH -43 0.92 Ethanol -84 1.15 Except for NMF, all data are taken from reference 177. b. 193 its analogs and 0221, the ligands are partially solvated by NMF ‘molecules. This makes the Cs+/ligand complexation formation constants in NMF are about ten times smaller than those in DMF. Dibenzo-24-crown-8, dibenzo-27-crown-9 and cryptand 0222 are too big to form complexes with NMF, so complexation reactions between cesium ions and these ligands are dominant. For these ligands there is little difference between the formation constants in NMF and in DMF. D. EEQQEflIIQ!§_£QB_IQBI§EB_§IHDIB§ 1) Create a new program which can solve the formation of 1:1, 2:1 complexes and collisional ion pairs. This would help us understand the influence of collisional ion pairs on the complexation reaction. 2) Perform complexation studies on lithium, sodium and thallium ions with crown ethers and cryptands. 3) Investigate the interactions. between. ligands and NMF, FA and DMF molecules via H-1 and C-13 NMR spectroscopy. W 194 W DETERNINATION OP ION PAIR PORMATION CONSTANTS BY TN! NMR TECHNIQUE: DESCRIPTION OF THE COMPUTER PROGRAN KINPIT AND SUBROUTINE EQUATION The equilibrium for ion pair formation can be expressed as M +X =MX (1.1) and Kip = [M+X-] / [14+][X-1v: = Kc / v: (L2) in which Kip’ K and v+are the thermodynamic ion pair 0 _ formation constant, the concentration equilibrium constant and the mean ionic activity coefficient respectively. By using the well known Debye-H'Lickel equation, v+ can thus calculated as follows: -logY:= [(4.198x106) z+z_ 11/2/(0'r)3/2] / [1 + (5.029x109a11/2)/(DT)1/2]. (1.3) In this equation 2+, z_ are the charges of the ions, I is the molar ionic strength which is 1/2 (iciziz) (C = 195 concentration summed over all species in the solution). D is the dielectric constant of the solvent and T and a are the temperature (K) and the closet distance of approach of the ions inK. The observed chemical shift is a population average of those of the free ion and the ion pair: i+e+ )X (1.4) aobs =6fo + 5ipxip = (‘f - lip f + ‘ip' Where X a [M+] / Ct; and C is the total concentration of f t metal ions in the system. Material balance gives 0t = [M+] + [M+x'] = [M+] + KC[M+]2. (1.5) Therefore +_._ 1/2 [M ] - [ l + (1 + 4KcCt) ] / (ZKC) (I.6) e + a- 1/2 xf [M 1 / 0t [ 1 + (1 + 4cht) 1 / (2Kc)0t(1.7) , 2 and Kc ..-Kip V4,. so that, finally sob, = [-1 + (1 + 4K1PV:C)1/21<6f - sip/(wig?) +5 (1.8) ip’ Three constants and two unknowns are used in the FORTRAN 196 code: COSNST (1) = CONST (2) = (01)“2 CONST(3) = a 5 f The two input variables are the concentration of the salt Ct and the observed chemical shift a obs which are designated as XX(1) and XX(2) respectively in FORTRAN code. Starting with an estimated values of 6 i and K i the p p' program fits the calculated chemical shift observed values by an iteration method. The SUBROUTINE EQUATION is listed on the following pages. 0000000 0000000 00000000 197 SUBROUTINE EQN IMPLICIT REAL*8(A-H,O-Z) IMPLICIT INTEGER*4(I-N) INCLUDE 'KINFIT:KINFITCOM.FOR/LIST' INCLUDE 'KINFIT:KINEQNCOM.FOR/LIST' Entry/Control Point GOTO (2,3,U,S,1,7,8,9,10,11,12,13) ITYPE ITYPE = 5 Initial call. No input has taken place 1 CONTINUE RETURN ITYPE = 6 Control card #1 and CONST have been input 7 CONTINUE NOUNK = 2 ! Set the dimensions of the problem NOVAR = 2 RETURN ITYPE = 3: 1 Experimental data has been read 8 CONTINUE RETURN ITYPE = 1: Evaluate algebraic equation and residual 000 00 35 00000000 0000000 CONTINUE 03MA=3XP(-4198000.0*SQRT(XX(1))/(C0NST(2)**3* (1.0+50.29*00NST(3)*SQRT(xx(1))/CONST(2)))) CONK=U(2)*(GEMA"2.0) 3:4.0*00NK*XX(1) S:((-1.0+SQRT(1.+F))/(2.*CONK*XX(1)))*(CONST(1)- U(1))+U(1) IF (IMETH .NE. -1) 0030 35 Simulations only RETURN Fits CONTINUE BRESID:S-XX(2) RETURN ITYPE = 2: Set the initial conditions for differential eqn's CONTINUE Y(1) = 1.0E-20 NOEQN = 1 ! Set the nuumber of equations RETURN °CONTINUE IF (0(1) .LE. 0.0) U(1) : ABS(U(1)) CONC1 = CONST(1) CONC2 = CONST(2) DY(1) = 0(1) ' (CONC1 - Y(1))*(CONCZ-Y(1)) -(U(1)/CONST(3))*Y(1) RETURN ITYPE = 4° Calculate the residual for differential eqn's , 0000000 0000 0000000 00000000 00000000 199 CONTINUE IF (IMETH .NE. -1) GOTO 20 CONTINUE CONST(6) is the molar absorptivity -- substitute exp value if known. CONST(6) = 5000.0 RESID : Y(1)+DLOG10(XX(2)/CONST(H))/(CONST(6)*0.2) RETURN ITYPE = 8: Calculate X(KVAR,I) the IPLT = 2 plotting mode RETURN -ITYPE = 9: FOP(I) : X(KVAR+1,I) for the IPLT : 3 -- mode CONTINUE RETURN ITYPE = 10: FOP(I) : X(KVAR+2,I) for the IPLT = 4 mode 200 11 CONTINUE ' RETURN ITYPE = 11: FU(I) <<<< x-axis; FO(I) <<<< yaxis (IPLT = 5) 00000000 12 CONTINUE RETURN 0000000 H .—3 l-< "U I!) II N 0 ‘D p—a H (D D. ID "9 n (D "I (D p. B C p—o W ('1‘ H0 O D 13 CONTINUE RETURN END 201 DETERMINATION OP COMPLEX FORMATION CONSTANTS BY THE NMR TECHNIQUE: DESCRIPTION OP THE COMPUTER PROGRAM KINPIT AND SUBROUTINE EQUATION A. CO PO 1' In the complexation reaction. between ligand L and metal ion M+ in solution, the equilibrium for a 1:1 complexation reaction can be expressed as: M + L = ML . (II.1) In the same way as described in Appendix I, the observed chemical shift can be expressed by: *' 2 2 2 2 2 d Obs [(KCM KCL 1) + (K CL + K CM - 2K CLCM + ZKCL fl, + 2x0" + 1)1/2](aM -6ML) / 2K0M +5“, (11.2) where K is the formation constant, CM is the total concentration of metal ions, CL is the total concentration of ligands, 5 ML is the chemical shift of complexes. In order to fit this equation, two constants and two unknowns are used in the FORTRAN code: 202 0(1) 0(2) = K ="ML cons'r (1) =0M CONST (2) =5" The two input variables are the concentration of the ligand CL and the observed chemlcla shift 5 obs which are designated ae XX(1) and XX(2) respectively in the FORTRAN code. Starting with an estimated values of K and ‘ML' the program fits the calculated chemical shift to the observed values by an iterative method. The SUBROUTINE EQUATION is listed on the following pages. 0000000 0000000 0000000 00000000 000000 203 SUBROUTINE EQN IMPLICIT REAL*8(A-H,0-Z) IMPLICIT INTEGER*N(I-N) INCLUDE 'KINFIT:KINFITCOM.FOR/LIST' INCLUDE 'KINFIT:KINEQNCOM.FOR/LIST' Entry/Control Point 0010 (2,3,u,5,1,7,8,9,10,11,12,13) ITYPE ITYPE = 5 Initial call. No input has taken place 1 CONTINUE WRITE (LUNOUT,8500) 1 Log the ID of this routine 8500 FORMAT (‘ *** EON: 1:1 Complex (16-JUN-86) ***') RETURN ITYPE : 6 Control card #1 and CONST have been input 7 CONTINUE NOUNK = 2 ! Set the dimensions of the problem NOVAR = 2 RETURN ITYPE = 3: ! Experimental data has been read 8 CONTINUE RETURN 204 2 ' CONTINUE IF(U(2).GT.O.) GO TO 1000 U(2)=1000 1000 CONTINUE A: U(2)*CONST(1) B: U(2)'XX(1) ‘ C= (CONST(2)- U(1))/(2. 'A) D: (B- A+1. )"2 CALC=((A-B-1.)+SQRT(D+H.*A))*C+U(1) IF (IMETH .NE. -1) GOTO 35 C C Simulations only C RETURN C c Fits C 35 CONTINUE RESID=CALC-XX(2) RETURN C c .............................. C C ITYPE = 2: Set the initial conditions for C differential eqn's C c .............................. C 3 CONTINUE RETURN C C .............................. C C ITYPE : 3 Evaluate the differential eqn 3 C c .............................. C 4 CONTINUE RETURN C ’ - C .............................. C C ITYPE : 4 Calculate the residual for differential C eqn's C C .............................. C 5 CONTINUE IF (IMETH .NE. -1) GOTO 20 Simulations only 000 RETURN 205 X(KVAR+2,I) for the IPLT : 4 F0(I) <<<< yaxis C C Fits C 20 CONTINUE C C CONST(6) is the molar absorptivity -- substitute exp C value if known. C CONST(6) = 5000.0 RESID : Y(1)+DLOG10(XX(2)/CONST(N))/(CONST(6)*0.2) RETURN C c .............................. C C ITYPE = 8: Calculate X(KVAR,I) the IPLT = 2 plotting C mode C c .............................. C 9 CONTINUE RETURN C c .............................. C C ITYPE = 9 FOP(I) : X(KVAR+1,I) for the IPLT = 3 C mode C c .............................. C 10 CONTINUE RETURN C c .............................. C C ITYPE : 10: FOP(I) : C mode C c .............................. C 11 ,CONTINUE RETURN C C .............................. C C ITYPE = 11 FU(I) <<<< x-axis; c (IPLT = 5) C C .............................. C 12 CONTINUE RETURN C c .............................. C 206 C ITYPE : 12: Called after simulation C . c .............................. C 13 CONTINUE RETURN END 207 B. , ON 0 S 0 ~ : QQNBLEZEE The equilibria for this reaction can be expressed as + + M + L = ML K = [ML 1 / [M 1[L1 (11.3) ML+ + L = ML K = [ML2+] / [ML+][L]. (11.4) Let [M+].T and [L]T denote the total concentrations of the metal ions and ligands respectively. Then + + + + [M 1T = [M ] + [ML 1 + [MLZ ] = [M+](l + K [L] + K K [L]2) (11 5) 1 1 2 ° + + [L1T = [L] 4» [ML 1 + 2[ML2 1 [L1(1 + K1[M+] + 2K1K2[M+][L]). (11.6) (II.5) and (II.6) can be arranged as + _ + 2 [M 1 - [M 1T / (1 + K1[L] + K1K2[L] ) (11.7) 2K K [M+][L]2 + [L](l + K [M+]) - [L1 = 0 (11 3) 1 2 1 T ' ' Therefore 208 + +2 + 12 [L] = {-(1 + 311M 1) + [<1 + K1 [M 1 + 8K1K21M ][L]TJ / 1 + / (4K1K2[M 1) (11.9) The observed chemical shift is: 5 obs = XM‘5M + XML‘ML + XML21’ML2 (“'10) In order to fit the calculated result with the experimental data, an expression for the relative mole fraction of all three species in terms of [M],r and [L]T is required. Two constants and four unknowns are used in the FORTRAN code: 5 U(4) U(l) = K U(2) = K 0(3) 1 2 ML 3 a ML2 + CONST (1) = [M 1T CONST (2) = ‘M The two input ‘variables are the total concentration. of ligand [L]T and the observed chemical shift ‘obs which are designated as XX(1) and XX(2) respectively. The SUBROUTINE EQUATION is listed on the folowwing pages. 0000000000 0 0 0 0 00 0 00 000000 0000000 00 0 00000 209 SUBROUTINE EQN This EON deals with the case of two competing equilibria. M+ + L = ML+ K1 ML+ + L = ML; K2 Unknowns: U(1) Formation constant, K1, for ML+. U(2) Formation constant, K2, for MLE. 0(3) Limiting chemical shift for ML+ U(4) Limiting chemical shift for MLE. Variables: XX(1) Concentration of ligand, (L). XX(2) Observed chemical shift of M nuclei. Constants: CONST(1) Total concentration of metal, M. CONST(Z) Chemical Shift for free metal. .----------------------------------------------“----‘---. .--------------------------------------------------------- IMPLICIT REAL*8(A-H,O-Z) IMPLICIT INTEGER'4(I-N) INCLUDE 'KINFIT:KINFITCOH.FOR/LIST' INCLUDE 'KINFIT:KINEQNCOM.FOR/LIST' INTEGER.“ NUMITR ! Maximum iterations of metal series ”DATA NUMITR/1000/ 11133 = 5: Initial call. No input has taken place 0000000 0000000 0000 0000000000 CONTINUE WRITE (LUNOUT,8500) ! Log the ID of this routine FORMAT (' *** EON: SANDCP { Sandwich Complexation', '26-SEP-86) } ***') RETURN CONTINUE NOUNK = 4 ! Set the dimensions of the problem NOVAR : 2 RETURN CONTINUE RETURN -This subroutine will solve a general polynominal The polynominal must be set so that the constant is positive. U(1) = ABS(U(1)) The Single letter variables that follow are the coefficients for the polynominal that is to be solved A=U(1)*U(2) A1=-A B = 2.0”A1'CONST(1) + A'XX(1) - U(1) 000 00 2005 200” 2003 2000 3000 3001 2001 2002 2006 999 2008 211 C = U(1)*XX(1) - U(1)'CONST(1) - 1 IF(XX(1).EQ.0.0) GOTO 2008 FREELi is the current value of the ligand being tested to see if it is the true free ligand concentration. VALUEi is the value of the polynominal FREELO = 0.0 FREEL1 = 0.00 VALUE1=XX(1) FREEL2=XX(1)/1.0E+12 CONTINUE VALUE2=((A1*FREEL2+B)‘FREEL2+C)*FREEL2+XX(1) IF(VALUE2.EQ.0.0) 0010 2001 IF(VALUE2.GT.0.0) 0010 2000 CONTINUE FREEL3=(FREEL2-FREEL1)/(VALUE1-VALUE2)*VALUE1+FREEL1 IF(ABS((FREEL3-FREELO)/FREEL3).LT.0.0000001) 0010 2002 VALUE3=((A1'FREEL3+B)*FREEL3+C)“FREEL3+XX(1) IF(ABS(VALUE3).LT.0.00000001) 0010 2002 IF(VALUE3.GT.0.0) 0010 2003 VALUE2=VALUE3 FREEL2=FREEL3 FREELO=FREEL3 0010 2004 CONTINUE VALUE1:VALUEB FREEL1:FREEL3 FREELO=FREEL3 0010 2004 CONTINUE VALUE1=VALUE2 FREEL12FREEL2 IF(FREEL2.LT.XX(JJ/100.) 0010 3000 FREEL2=FREEL2+XX(1)/100. 0010 3001 CONTINUE FREEL2:FREEL2*10. 'CONTINUE IF(FREEL1.GT.XX(1)) 0010 2006 0010 2005 CONTINUE FREEL=FREEL2 0010 2007 CONTINUE EREEL=EREEL3 0010 2007 CONTINUE IF(LAP.NE.3) 0010 2008 NRITE (LUNOU1,999) EORMAT(/,5x,29au NO NEGATIVE VALUE FOUND **) CONTINUE 3002 000 000 35 00000000 0000000 212 FREEL=0.0 CONTINUE EREEM = CONST(1)/ (1.0 + U(1)*FREEL + U(1)'U(2)*EREEL*FREEL) FREEML - U(1)*FREEM*FREEL FREMLL - U(2)”?REEML'FREEL CALC = ( FREEH'CONST(2) + FREEHL*U(3) + FREMLL*U(H))/ CONST(1) CONTINUE IF (IMETH .NE. -1) GOTO 35 Simulations only RETURN Fits CONTINUE RESID:CALC-XX(2) RETURN ITYPE = 2: Set the initial conditions for differential eqn's CONTINUE RETURN CONTINUE RETURN ITYPE = fl: Calculate the residual for differential eqn's 213 C .- 5 ' CONTINUE IF (IMETH .NE. -1) GOTO 20 C C Simulations only C RETURN C c Fits C 20 CONTINUE RETURN C c .............................. C C ITYPE : 8: Calculate X(KVAR,I) the IPLT = 2 plotting C mode C c .............................. C 9 CONTINUE RETURN C c .............................. C C ITYPE : 9: FOP(I) : X(KVAR+1,I) for the IPLT = 3 C mode C c .............................. C 10 CONTINUE RETURN ITYPE : 10: FOP(I) mode X(KVAR+2,I) for the IPLT : 3 00000000 11 'CONTINUE RETURN ITYPE = 11: FU(I) (<<< x-axis; F0(I) <<<< yaxis (IPLT = 5) 00000000 12 CONTINUE RETURN c .............................. 00000 21H ITYPE : 12: Called after simulation 13 CONTINUE RETURN END 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 215 w. Ostwald, Z1 £0!§ik;.§h§mi. 2, 270 (1888). F. Kohlrausch and L. Holborn, "Das Leitverm gen der Elektrolyte", Teuber, Leipzig (1916). L. Onsager, Z; Phx§,, 26, 277 (1927). L. Onsager and R. M. Fuoss, l1 Bugs, Qhem6, 66, 2689 (1932). R.M. Fuoss and C. A. Kraus, 66 Am, gngm6 6666, 66, 476 (1933). ' L. Onsager and R.M. Fuoss, 66 23266 gh6m6, 61, 668 (1957). R.F. Prini, 116366 Eazaday §QC., 66, 3311 (1969). E. Pitt. 2:991 8221 S921 Ar. 211. 43 (1953)- R.M. Fuoss and K. L. Hsia. 2:921 £2112 82292 $911. 52. 1550 (1967). J.C. Justice, 16 636m; Ehxg6, 66, 353 (1968). J.C. Justice, ELQQSIQQDLEE Acta., 16, 701 (1971). R.M. Fuoss, 16 Eth6 thmi. 12, 525 (1975). T._Arnold and M. E. Packard, g6 chem, Enxg6, 12, 507,(1951). J.F. Hinton and K. R. Metz, " NMR of Newly Accessible Nuclei", Volume 2, Academic Press, New York, Chapter 14 (1933). E. G. Bloor and R. G. Kidd, 6636 16 gngm6, 56, 3425 (1968). E. G. Bloor and R. G. Kidd, 6636 Q6 Qngm6, 66, 3926 (1972). C. Deverell and R. E. Richards, M211 Ehng, 16, 551 (1966). 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 216 J.D. Halliday, R.E. Richards and R.R. Sharp, 2199; Boy, 9991 Landau. 3135. 45 (1969)- C. Hall, R.E. Richards and R. R. Sharp, 2:221 8921 §OQ. Landgn. 111A. 297 (1974)- A-I- POPOV. Burs 82211 Qnsm1. i1. 101 (1979)- R. H. Erlich and A. I. Popov, is An; Qngm6 $666, 22, 5620 (1971). W.J. DeWitte, L. Liu, E. Mei, J.L. Dye and A.I. Popov, 11 591n1 Chem1. 5. 337 (1977)- S. Khazaeli, A.I. Popov and J. L. Dye, 11 Eh231 gngm6, 95. 4238 (1982). W.J. Dewitte and A. I. Popov, 21.52101 gngm6, 6, 231 (1976). J.F. Hinton and K. R. Metz, 26 fian; Qth6, 2, 197 (1930). H.G. Hertz, V. Versmold and C. Yoon, 26:6 Bungen; £h2§1 thm1. El. 577 (1933)- H. G. Hertz and R. Mazitov, 53:1 Bunssn1 Rhys; thm1, 2:, 1103 (1981). C. Deverell, M211 £h¥§1p 16, 491 (1969). H. G. Hertz, R. Tutsh and H. Versmold, 36:6 finnggn; RDXEL 9h9m1. 15. 1177 (1971)- H.G. Hertz, H. Holz, G. Keler, H. Versmold and C. Yoon, 59:1 2909901 20221 thm1. 19. 493 (1974)- H. G. Hertz, G. Stalidis and H. Versmold, is thm; Bh¥§1: §§. 177 (1959)- H2.Weing§rtner and H. G. Hertz, 26:6 231116611L BBXQL Qhfils. 51. 1204 (1977)- J.C. Evans and G. Y.-S. Lo, 11 Eh¥§1 Qn6m6, 62, 3223 (1965). W.F. Edgell, A. T. Watts, J. Lyford and W. H. Risen, 16 Am; thm1 £991. 22. 1855 (1966). w.r. Edgell, J. Lyford, R. Wright, W. M. Risen and A.T. Watts. 11 Ami thm1 5991. 22. 2240 (1970). B.w. Haxey and A. I. Popov, 11 Am; Chem; £921. 21, 20 (1969). 41 49 50 51. 520 53. 54. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 217 J. L. Wuepper and A. I. Popov, 26 Am; thm; 2666, 21, 4352 (1969). A- 1- Popov, Burs 82211 98281. 11. 275 (1975). A.T. Tsatsas, R.H. Stearns and W.M. Risen, 26 Am; ghgm6 5991. 25, 5247 (1972). Y.M. Cahen and A. I. Popov, 21 26166163 6h6m6, 6, 599 (1975). W.F. Edgell, A. T. Watts, J. Lyford and W. M. Risen, 16 Am1 Chem1 £921. 88. 1855 (1966)- W.F. Edgell, J. Lyford, R. Wright, w. M. Risen and A.T. WattS. 21 881 98281 5991. 22. 2240 (1970)- W.F. Edgell, J. Lyford, A. Barbetta and C. I. Lose, 26 An1 §h§m1 £921. 21. 6403 (1971)- W.F. Edgell and J. Lyford, l1 Am6 Cbsm1.§991. 22, 6407 (1971). W.P. Edgell and A. Barbetta, 11 Am; Qhfimi £221. 96, 415 (1974). J. B. Gill. BRIE ABEL; §h§m1p.ii. 1355 (1931)- D.N. Waters, " Vibrational Spectra of Electrolyte Solutions "in pecialist Periodical Reports," Molecular Spectroscopy" Vol. 6, The Chemical Society, London (1979). A.G. Miller and J. W. Macklin, l1 Rhys, gngm6, 62, 1193 (1985). R.H. Atherton and S. I. Weisman, J; Am; thm; $221. 61, 1330 (1961). P.J. Zandstra and S. I. Weisman, 16 Am; §h6m6 $666, 61, 4408 (1962). K. Hételmann, J. Jagur-Grodzinski and M. Szwarc, J6 AmL Chan; £291. 21. 4545 (1959)- H. Smith and H. C. R. Symons, IIQn§1 {graggy £921. 61, 338 (1957). T.R. Griffiths and R. H. Wijayanayake, 116n§6 {gigggy £991. 55. 1563 (1970). T.R. Griffiths and R.H. Wijayanayake, £1£1§1 EQIQQQ! 1, 62, 1899 (1973). 6. 69 70 71. 72. 73. .74... 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 218 W. R. Gilkerson and A. M. Roberts, 11 221 63661 §oc,, 192. 5181 (1980). W.R. Gilkerson and K. L. Kendrick, 11 23261 66631, 26, 5352 (1984). K.L. Kendrick and W. R. Gilkerson, 21 661n1 gngm1, 16, 253 (1986). ' C.J. Pedersen, 11 Am1 CREE; 6661, 62, 7017 (1967). C.J. Pedersen, 11 Am1 gngm1 6661, 22, 391 (1971). C.J. Pedersen, 11 6:61 gngm1, 66, 254 (1971). B. Dietrich, J. H. Lehn and J. P. Sauvage, Iggrgnggzgn 19111. 2885 (1969). E. Graf and J. M. Lehn, 21 561 CREE; 6661, 21, 5022 (1975). J-M- Lehn. 8921 98981 3921. 11. 49 (1978)- J.H. Lehn, E. Souveaux and A. K. Willard, 61 Am1 thm; £221. 129: 4915 (1973)- R.H. Izatt, J. S. Bradshaw, s.A. Nielsen, J.D. Lamb, D. Sen and J.J. Christensen, £82m; 3621, 66, 271 (1985). R.H. Izatt and D.P. Nelson, 6616366, 165, 443 (1969). P. Seiler, H. Dobler and J. D. Dunitz, 56;; 911518119911. 125, 2744 (1974). R.H. Izatt, R. E. Terry, D. P. Nelson, Y._Chan, D. J. Eatough, J.S. Bradshaw, L. D. Hansen and J. J. Christensen, 21 Am1 QhfiE; 6661, 26, 7626 (1976). D.E. Fenton, H. Mercer, N. S. Poonia and H.R. Truters, £89m; 9988281. 66 (1972)- H.-A. Bush and H. R. Truter, 11 gngm1 6661, Egrkin 11, 345 (1972). E. Hei, A.I. Popov and J.L. Dye, 11 261 thm1 6691, 22, 6532 (1977). R. D. Boss and A. I. Popov, 136161 ghgm1, 26, 1747 (1986). J.J. Christensen, D. J. Eatough and R. H. Izatt, Chem; 3321, 15, 351 (1974). J.H. Lehn and J.P. Sauvage, 11 561 thm1 6661, 21, 6700 81 87 88 89. 90. 910 92. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 89. 90. 91. 92. 219 (1975). R.H. Izatt, D. J. Eatough and J.J. Christensen, 6616661 2989189. 15. 161 (1973)- , H.F. Frensdorff, 11 Am1 §h6m1 $991..21. 600 (1971). D.J. Cram, R. C. Helgeson, L. R. Sousa, J. M. Timko, M. Newcomb, P. Moreau, F. Dejong, G. W. Gokel, D. H. Hoffnam, L. A. Domeier, S. C. Peacock, K. M. Madan and L- Kaplan. Burg 8 89211 98981. 11. 327 (1975)- E. Mei, A.I. Popov and J. L. Dye, 11 thg; thm;. 61. 1877 (1977). M. Shamsipur and A. I. Popov, 11 Am1 thm; £991. 121. 4051 (1979). E. Mei, L. Liu, J. L. Dye and A. I. Popov, 61 661n1 QDEEL. 9. 771 (1977)- A.J. Smetana and A. I. Popov, 11 661n1 gn6m1, 2, 183 (1980). M. Shamsipur, G. Rounaghi and A. I. Popov, 11 661n1 98981. 2. 701 (1980)- J.D. Lin and A. I. Popov, 11 261 63661 6691, 122, 3773 (1981). P. A. Mosier-Boss and A. I. Popov, 11 Am1 thm1 629;, 121. 6168 (1985). E. Kauffman, J. M. Lehn and J. P. Sauvage, 36121 ERIE; AQSQ1. 52. 1099 (1975)- D. K. Cabbiness and D. W. Margerum, 11 Am1 SEEM; 6921, 21, 6540 (1969). R. D. Hancock and A. E. Martell, ngmgntg 6n Ingrg; m: Q: 237 (1988)“ J.A. Riddick and W.R. Hunger, "Techniques of Chemistry, II, Organic Solvents", Wiley-Interscience, New York (1970). R-L- JoneS. 11.M91L'§292312§£1. 2. 581 (1958)- T. Miyazawa, T. Shimanouchi and S. Mizushima, Qngm1 ERXEE. 21. 408 (1955)- I- Suzuki. 59112.§hng £921 £281. 35. 540 (1962)- R.A. Russel and H.W. Thompson, 622622663161 56661, 2, 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 220 138 (1956). T. Miyazawa, 11 M211 6266616661, 1, 155 (1960). R.A. Elzaro, 61661 522:1; Int; B. 11. 52111 ghgmL £291 Inn21‘fill 1052 (1973). M. Kitano and K. Kuchitsu. 631 (1974). L.A. LaPlanche and M.T. 337 (1964). RogerS. 11 881 989m; 5991. 85. M.J. Aroney, R.J.W. £991, 1, 3179 (1965). Y. Sugawara, A.I. Hirakawa, M. Tsuboi, S. Kato and K. Morokuma, CREE; 2n261, 62, 339 (1981). H. Ohtaki, S. Itoh and B. M. Rode, 66111 thm; 6921 innl. 52. 271 (1985)- V.A. DurOV. 39581 11 £8251 98982. 25. Le Fevre and A.N. Singh, 11 thm1 1616 (1981). G.R. Leader and J. F. Gormley, 11.5m1.§h3n1.5991. 11, 5731 (1951) S.J. Bass, W. I. Nathan, R.H. Meighan and R.H. Cole, 11 20281 dhsm1. 58. 509 (1964)- P. Winsor and R. H. Cole, 61 26261 6n6n1, 66, 2486 (1982). P-P- Rastogi. £3112 98981 $991 inn2. $3. 2442 (1970)- P-P- Rastogi. 2:291 Indian N2111 £211 Adddl. Eart_8. 51, 345 (1981). J. Bukowska, 11 8611,6636661, 26, 1 (1983). 0. D. Bonner and C. F. Jordan, 26261611 gngm1 8 2626166, 6, 293 (1976). R.C. Paul, J. P. singla, M. S. Lamba, D.S. Gill and.P. Narnia. Indian 11 98981. 11. 1024 (1973)- L. Weeda and G. Somsen, 3661 1:621 gn1m1, 66, 159 (1966). L. Weeda and G. Somsen, 3661 {11621 66161, 66, 263 (1967). C.H. Criss, R.P. Held and E. Luksha, 12, 2970 (1968). l1.2h¥§2.§h2n;r 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 221 D.S. Gill, J. P. Singla, R.C. Paul and S.P. Narula, .11 Chani 2991. 522 (1972)- J. Emsley, D.J. Jones, J.M. Miller, R.E. Overill and R.A. Waddilove, 61 An1 Qngm1 6661, 162, 24 (1981). C. Carmona, G. Eaton and M.C.R. Symons, 11 gngm1 £291. snani damnunI. 873 (1987)- R.C. Paul, D.S. Gill, J.P. Singla and S.P. Narula, Indian 11 Chami. 2. 63 (1971)- T. Pontani and B. M. Rode, 36n662h1 thm1, 166, 1153 (1977). T- Shedlovsky. II Eranklin Insti. 222. 739 (1938)- R-D- Singh and R- Gopal. 22111 Chani Sada Inna. 42. 2088 (1972). C.M. French and K. H. Glover, 116n§1 Kazaflax1 $221. 51. 1418 (1955). G.P. Johari and P. H. Tewari, 11 BREE; Qn6m1, 62, 3167 (1965). R.P. Held and C. M. Criss, 61 23261 gngm1, 62, 2611 (1965). E. Lnksha and C.M. Criss, 21 2ny§1 gngn1, 16, 1496 (1966). I. S. Perelygin, S. V. Izosimova and Yu. M. Kessler, znnri Strnki KhinI. 2. 390 (1968). R.W. Briggs, K.R. Metz and J.P. gnani. 2. 479 (1979)- H. Weingfirtner. II 20251 Qnani. (1982). I'Io'WBi-ng'a'l’tner. it. 91191.. 5.9.9.2.: m m .1! 18.! 3053 (1982). Hinton. II Saintian (Wiesbaden), 166, 23 M. Q. Martin, G. J. Martin and J. "Practical NMR Spectroscopy", London. (1980) -J. Delpeuch, Heyden and Sons, Ltd., D.N. Bhattacharyya, C. L. Lee, J. Smid and M. Szwarc, 11 £8221 98281. 22. 608 (1965)- A. Rykowski and H. L. Vanderplas, £221. 21. 204 (1975). £991 11 8921 Natni 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. 147. .F.L. Cook. 11 9:91 ChanI. 32. 222 G.W. Gokel, D. J. Cram, C. L. Liotta, H. P. Harris and 2445 (1974). B.L. Haymore, J. D. Lamb, R. M. Izatt and J. J. Christensen, 1ngzg1 gngm1, 21, 1598 (1982). C. Brevard and P. Granger, " Handbook of High Resolution Multinuclear NMR", John Wiley and Sons (1981). J. C. Lindon and A. G. Ferridge, 21291 in NMR 2229119221. 11. 27 (1980)- G.C. Templeman and A. L. Van Geet, 61 Am1 Qh6m1 6661, 21, 5578 (1970). R. D. Boss, Ph.D. Thesis, Michigan State Univrsity, East Lansing, Michigan (1985). J. L. Dye and 443 (1971). v. A. N1celY1 1.7. gm; 351119.... 5.8.: H.-H. Nam, Private Communication. J- 8- Shih and A- I- POPOV. Inardi 82911 Chani Latti. 11, 105 (1977). W. J. Dewitte, R. C. Schoening and A. I. Popov, 136:61 HEEL; £83m; Lfitta. 12. 251 (1975)- R. W. Briggs and J. F. Hinton, 21 Eana gngm1, 6, 827 (1977). Y. M. Cahen, P.R. Handy, E.T. Roach and A.I. Popov, 11 BREE; Shana. 12. 30 (1975) J- W- Akitt. II QbanI.2991 Iaradax Iranai 1. 11. 1557 (1975). C. Deverell, Bragri NEE 2222222221. 1. 278 (1969)- J..Kondo and J. Yamishita, 21 Rhys. SEER; §Qli§§a.19. 245 (1959). D. Ikenberry and T.P. Das, 11.6ngm1 Enyg1, 16, 2199 (1965). D. Ikenberry and T.P. Das, 11 6n6m1 2h261, 66, 1361 (1966). A.K. Covington and K. E. Newman, In: Modern Aspects of Electrochemistry, Vol. 12, Eds. J.D. Bockris and B.E. Conway, Plenum Press, New York (1977). 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163. 164. 165. 223 N- Bjerrum. Kali Dani Yidanski Salaki Matizzxai Maddi. ~1, No. 9 (1926). A. I. Mishustin and Y. M. Kessler, 61 691n1 66661, 1, 779 (1975). J. W. Akitt, " Multinuclear NMR ", Press (1987). ' T.P. Young, Y.C. Wu and A.A. Krawetz, 01229221 Egzgdax £921. 21. 37 (1957)- Y. c. Wu, N. B. Smith and T. P. Young, 11 BREE; gngm1, Chapter 7, Plenum 62, 1868 (1965). J.H. Stern and G.W. Anderson, 11 £hx§1 gn6m1, 66, 2528 (1964). R.H. Wood and R. W. Smith, 61,2h261 gngm1, 62, 2974 (1965). M.S. Stakhanova, V. Vasilev and Yu. A. Epikhin, 36661 11 beaI 98:81. 21. 182 (1953)- C. Deverell and R. E. Richards, u611 EDYRa. 16, 421 (1969). 6- Kabisch and M. Klose. 21 28221.9bania. Leipzig. 266. V.I. Baranovskil, T.T. Ermolaeva, V. I. Chizhik, 0. V. Sizova and N.V. Ivanova, 6621 11 gngm1 £h261, 6, 861 (1987). I.S. Perelygin, 6661 11 6663661, 13, 360 (1962). I. S. Pominov and A. z. Gadzhiev, 1121,261611 61113:, No.6, 19 (1965). I. S. Pominov and T. E. Pavlova, zngrngl 2213166661 SRQKEIQEBQRIII. 11. 277 (1959)- J.A; Pople, W. G. Schneider. and H. J. Bernstein, High -resolution Nuclear Magnetic Resonance, Chap. 15. (New York, McGraw-Hill Book Co.) (1959). 687 (1985). C. E. Schafter and C. K. Jorgensen, 61 1n6161 H3211 Shani. 2. 143 (1958)- Y.-C. Lee, Ph.D. Thesis, Michigan State University, East Lansing, Michigan (1983). A. Ellaboudy and J. L. Dye, 11.881 Shani 292;. 125. 6490 (1983). 166. 167. 168. 169 170. 171. 172. 173. 174. 175. 176. 177. 224 S. Mooibroek, R. E. Wasylishen, R. Dickson, G. Facey _and B.A. Pettitt, 11 M661 366661, 66, S42 (1986). M. S. Greenberg and A.I. Popov, 11 661n1 66661, 6, 653 (1976). C.J. Pedersen, 11 231 ghgm1 6661, 22, 386 (1970). W. H. Watson, J. Galloy, D. A. Grossie, E. vagtle and W.M. Miller, 61 6:61 6h6m1, 162, 347 (1984). J-M- Lehn. Strnainra Banding. 1a. 1 (1973)- D.L. Johnston and W.D. Horrocks, 182191 thm;. 12. (1971). 687 R.W. Kluiber and G. Sasso, 226 (1970). Inargi China Astai. A. M. Herceg and R. 435 (1970). WeiSS. Indrsi 82911 dbani Latai. a. D. Live and S. I. Chen, 61 Am1 66661 6661, 26, 3769 (1976). 299 A. Hourdakis and A. I. Popov, 11(661n1 £hgm1. 6. (1977). N. Ahmad and M. C. (1977). ' Day, 21 Ana §h§m1 fiflfia. 22. 941 E. Schmidt, J.-M. Tremillon, J.-P. Kintzinger and A.I. POPOV. 11.8n1 98981 2991. 125. 7563 (1983)-